Next Article in Journal
Differential Gene Expression Induced by Different TLR Agonists in A549 Lung Epithelial Cells Is Modulated by CRISPR Activation of TLR10
Next Article in Special Issue
The Essentials on microRNA-Encoded Peptides from Plants to Animals
Previous Article in Journal
Bioactivity and In Silico Studies of Isoquinoline and Related Alkaloids as Promising Antiviral Agents: An Insight
Previous Article in Special Issue
Role of circRNA in E3 Modification under Human Disease
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Non-Coding RNAs in the Regulation of Hippocampal Neurogenesis and Potential Treatment Targets for Related Disorders

1
Department of Anatomy, Institute of Neurobiology, Medical School, Co-Innovation Center of Neuroregeneration, Nantong University, Nantong 226001, China
2
School of Pharmacy, Nantong University, Nantong 226001, China
3
Key Laboratory of Inflammation and Molecular Drug Target of Jiangsu Province, Nantong 226001, China
4
Central Lab, Yancheng Third People’s Hospital, The Sixth Affiliated Hospital of Nantong University, Yancheng 224001, China
*
Author to whom correspondence should be addressed.
Biomolecules 2023, 13(1), 18; https://doi.org/10.3390/biom13010018
Submission received: 14 November 2022 / Revised: 17 December 2022 / Accepted: 19 December 2022 / Published: 22 December 2022
(This article belongs to the Special Issue Circular RNAs: Functions, Applications and Prospects)

Abstract

:
Non-coding RNAs (ncRNAs), including miRNAs, lncRNAs, circRNAs, and piRNAs, do not encode proteins. Nonetheless, they have critical roles in a variety of cellular activities—such as development, neurogenesis, degeneration, and the response to injury to the nervous system—via protein translation, RNA splicing, gene activation, silencing, modifications, and editing; thus, they may serve as potential targets for disease treatment. The activity of adult neural stem cells (NSCs) in the subgranular zone of the hippocampal dentate gyrus critically influences hippocampal function, including learning, memory, and emotion. ncRNAs have been shown to be involved in the regulation of hippocampal neurogenesis, including proliferation, differentiation, and migration of NSCs and synapse formation. The interaction among ncRNAs is complex and diverse and has become a major topic within the life science. This review outlines advances in research on the roles of ncRNAs in modulating NSC bioactivity in the hippocampus and discusses their potential applications in the treatment of illnesses affecting the hippocampus.

1. Introduction

Non-coding RNAs (ncRNAs) are special RNA transcripts that compose a major portion of the human transcriptome; more than 90% of the human genome is actively transcribed, but only 2% of the entire genome encodes protein-coding RNAs [1,2]. With the exception of several ncRNAs with open reading frames, ncRNAs generally do not encode proteins. Nonetheless, they act as important regulators of development, proliferation, transcription, post-transcriptional modification, apoptosis, cell metabolism, and other biological processes [3]. Transfer RNAs (tRNAs), ribosomal RNA (rRNA), small nuclear RNAs (snRNAs), and ribozymes are all ncRNAs. The structure and function of these non-coding rRNAs have been well documented. Other types of ncRNA include microRNAs (miRNAs), piwi-derived small RNAs (piRNAs), small interfering RNAs (siRNAs), tRNA-derived small RNAs (tsRNAs), long non-coding RNAs (lncRNAs), small nucleolar RNAs (snoRNAs), circular RNAs (circRNAs), and pseudogenes (Ψ). These ncRNAs vary in size, ranging from small to large, as well as in spatial structure, and are usually classified on the basis of their length. Small ncRNAs are less than 200 nucleotides long and include siRNAs, miRNAs, and piRNAs, whereas transcripts longer than 200 nucleotides are termed lncRNAs. Small ncRNAs, such as miRNAs and piRNAs, are <32 nucleotides in length, whereas lncRNAs can be as long as 5000 nucleotides. These RNAs are involved in virtually all cellular processes. Additionally, lncRNAs also regulate gene and protein expression levels through diverse complex mechanisms.
ncRNAs are specifically expressed in different tissues and organs at different stages of development or disease [4]. Many ncRNAs have been found to control transcription and translation, thus ultimately affecting development and disease progression; however, the exact activities of most ncRNAs remain unclear. Moreover, recent studies have indicated that ncRNAs act as key regulators of cell proliferation and death [5]. In recent decades, research on the activity of neural stem cells (NSCs) has focused primarily on niche, molecular, and protein-coding genes. However, studies increasingly indicate that ncRNAs play key regulatory roles in NSC activity. Many factors, including ncRNAs in the adult hippocampus, have been found to affect neuronal differentiation and growth by modulating various signaling pathways [6].
For instance, the expression of miRNAs is highly concentrated in the brain tissues of both humans and rodents and is distributed in several anatomical areas of the brain. For example, hippocampus-specific expression of miR-128a/b, miR-218, miR-138, miR-222, miR-26a, miR-221, and let-7c has been reported [7]. miRNA expression varies by cell type. miR-124 and miR-128 are two examples of miRNAs that are expressed primarily in adult neurons but not in glial cells, whereas miR-23 is found in astrocytes [8] and miR-92b is specifically expressed in NSCs [9]. Furthermore, miR-125b and miR-93 are highly expressed in NSCs of the subventricular zone (SVZ) region [10]. miRNAs have also been implicated in adult neurogenesis.
The hippocampus is essential for complex neurological functions such as learning, memory, and the processing of emotional data. Adult resident NSCs in the subgranular zone (SGZ) undergo proliferation, migration, differentiation, and maturation into granular cells in the hippocampal granular cell layer under the influence of pathological stimuli. They also produce axons that target and form functional synaptic connections with neurons in the CA3 region, integrating into the functional neural loop of the hippocampus [11,12,13]. Hippocampal neurogenesis in the hippocampal dentate gyrus (DG) plays a prominent role in the formation and self-repair of spatial learning and cognitive memory [14,15].
ncRNAs have been increasingly observed participating in hippocampal neurogenesis and functional reconstruction, including NSC regeneration, neural circuit repair, and recovery of learning and memory. This review summarizes progress in the understanding of the roles of ncRNAs in the regulation of hippocampal neurogenesis, particularly the differentiation of NSCs, and describes potential applications in the treatment of hippocampus-associated diseases, thus providing directions for future research and clinical practice.

2. Effects of miRNAs on the Differentiation of Hippocampal NSCs

miRNAs are highly conserved ncRNAs approximately 22 nucleotides in length. They are ubiquitous in eukaryotes and have several biological characteristics, including (1) high conservation, (2) temporal expression specificity, and (3) tissue expression specificity. The nucleotide sequence at position 2−8 of the 5′ end of mature miRNA is generally agreed to be the seed sequence (the region where miRNAs specifically bind targeted mRNAs). The roles of miRNAs and their target mRNAs are not one-to-one. A single miRNA usually targets and regulates multiple mRNAs, and the same mRNA can also be targeted and regulated by multiple miRNAs, thus forming a complex overlapping regulatory network. As summarized in Table 1, miRNAs are believed to regulate the differentiation, proliferation, and apoptosis of NSCs.

2.1. miR-132

miR-132, an important miRNA affecting the differentiation of NSCs, is highly expressed in brain-specific DG regions and neurons and was originally observed in the nerve tissues of mice [16], with later observations in humans [17], zebrafish [18], and cows [19]. Studies showed that overexpression of miR-132 promoted the differentiation of NSCs into astrocytes but inhibited neuronal formation [20,21], and miR-132 knockdown impaired integration of newborn neurons into the adult dentate gyrus [22]. The level of miR-132 decreased in the hippocampus during progression of Alzheimer’s disease (AD); this miRNA is required for proliferation of neural precursors and neuronal differentiation in the adult dentate gyrus [23]. The above reports indicate that miR-132 is firmly involved in the progress of hippocampal neurogenesis.
Evidence implies that miR-132 regulates the differentiation of hippocampal NSCs through a variety of actions, particularly in early stages of neuronal development. In early stages of NSC differentiation, miR-132 fine-tunes Notch signaling (very important for the maintenance of undifferentiated NSCs [24]) through its target, Ctbp2, and regulatory pathways involving REST and Sirt1, which inhibit the proliferation of radial glial cells (early stage NSCs) and promote the differentiation of oligodendrocytes [25]. MeCP2 (X-linked methyl-CpG binding protein 2) is a multi-functional epigenetic factor modulating the microenvironment and is critical for normal brain development [26], including hippocampal neurogenesis [27]. MeCP2 is the target of and is regulated by miR132. When miR-132 is overexpressed, it causes simultaneous downregulation of MeCP2 and inversely blocks miR132 mediated-increases in MeCP2 [20,28,29]. miR132 itself was induced by BDNF, a neurotransmitter modulator whose expression was in turn controlled by MeCP2, while the loss of MeCP2 reduced BDNF and miR132 levels in vivo [29]. Obviously, MeCP2-BDNF-miR132 forms a regulatory feedback loop compensating the homeostasis regulatory network in the brain to ensure microenvironmental balance (as shown in Figure 1). ERK1/2 signaling is crucial for neuronal proliferation and differentiation, as well as neurite outgrowth. Moreover, blocking the ERK1/2 signaling pathway decreases miR-132 levels and neuronal differentiation. This signaling pathway therefore controls miR-132 production—a crucial step in the early neuronal lineage of stem cells [30]. The reports above signified the complicated mechanisms mediating the role of miR-132 controlling neurogenesis.

2.2. miR-199 and mir-214

Mellios and colleagues found that miR-199 and miR-214 are strongly upregulated in patients with Rett syndrome and MeCP2-deficient neuronal progenitor cells (NPCs), and also found that they are important effectors in early neuronal development through multiple effects on the AKT and ERK signaling pathways [31]. In MeCP2-deficient NPCs, upregulated miR-199 and miR-214 have different regulatory effects on the ERK and AKT signaling pathways. The upregulation of miR-214 decreases the expression of PTEN, whereas the upregulation of miR-199 significantly inhibits the expression of PAK4 in MeCP2-deficient neurons. In the hippocampus of MeCP2-deficient mice, miR-199/214 is upregulated and expression of targets PAK4 and PTEN decreases, thus resulting in differential regulation of upstream ERK and AKT signaling pathways in neurogenesis on the basis of inactivation of the ERK pathway and activation of the AKT pathway. Moreover, changes in the expression of miR-214/PTEN and miR-199/PAK4 in 3-week-differentiated neurons derived from iPSCs led to both ERK and AKT activation, thus indicating a compensatory effect on the activation of the two central molecular centers [31]. As shown in Figure 1, miR-132, MeCP2, miR-199/214, and the ERK and AKT pathways form a regulatory loop modulating the differentiation of hippocampal NSCs.

2.3. miR-140-5p

According to previous research, miR-140-5p is highly expressed in patients with post-stroke depression, and its expression positively correlates with depression severity [32]. miR-140-5p was likewise upregulated in a mouse model of bilateral common carotid artery ligation, and consequently may limit hippocampal neurogenesis and exacerbate cognitive impairment [33].
miR-140-5p has been confirmed to target and downregulate expression of Prospero Homeobox 1 (Prox1), which mediates the inhibition of hippocampal neurogenesis [33,34]. Prox1 is found in the thalamus, cortex, hypothalamus, DG, and cerebellum during the prenatal, postnatal, and adult development phases [35,36,37]. Under the influence of Wnt signaling, the development of nascent neurons in the granular layer of the adult hippocampus is regulated and controlled by Prox1 [38]. Additionally, Prox1 facilitates NSC differentiation and cell cycle withdrawal by inhibiting the Notch pathway [39].
Prox1 is essential in the differentiation and maturation of glutamatergic central neurons and intermediate progenitors during hippocampal neurogenesis in adults [38,40,41]. In addition, the staining of GFAP, DCX, and MAP-2 during miR-140-5p overexpression has indicated that miR-140-5p inhibits neuronal differentiation of NSCs but promotes astrocytic differentiation via the production of synapse-associated proteins. The ERK/MAPK signaling pathway is stimulated by miR-140-5p through negative regulation of the downstream target Prox1 [34]. In conclusion, miR-140-5p inhibited neuronal differentiation via negatively regulating Prox1 and activating the ERK/MAPK signaling pathway.

2.4. miR-137

miR-137 is enriched in both embryonic and adult brains and has been reported to promote adult SVZ NSC differentiation into neurons and to inhibit their maturation in the adult hippocampus [42,43]. The gene encoding this miRNA has been confirmed to be located on human chromosome 1p22 within the non-protein-coding gene BRAMY2014205 [44], and has also been observed regulating the expression of more than 1000 predicted target genes [45,46]. Therefore, identifying the cascades involving miR-137 and the fate of NSCs is of great importance.
Functional experiments have confirmed that miR-137 is involved in neuronal development, and studies in a mouse model revealed that miR-137 suppresses cellular proliferation while facilitating neurodevelopment. Forced expression of miR-137 in mouse embryonic NSCs decreases proliferation but enhances premature neuronal differentiation [42]. Moreover, NSCs within the SVZ display similar effects [43]. Recently, Channakkar et al. found that miR-137 overexpression in human iPSC-derived NSCs (hiNSCs) significantly increases newly formed neurons [47]. However, miR-137 exhibits different effects on adult hippocampal neurogenesis in the dentate gyrus. Overexpression of miR-137 disrupts neuronal differentiation as well as dendritic arborization [48]. Szulwach et al. also found that miR-137 overexpression inhibits maturation and promotes proliferation of adult NSCs, whereas antagonism of miR-137 promotes the differentiation of neurons and decreases proliferation [46]. This apparent disparity is probably due to the diverse ways in which miR-137 participates in various phases of neuronal development, as well as the inherent characteristics of cells in various neurogenic zones. These studies have demonstrated that correct neuronal differentiation, proliferation, and modulation of neurogenesis depends on balanced expression of miR-137.
The mechanism mediating the effects of miR-137 on the neuronal differentiation of NSCs may be related to the feed-forward self-regulating loop between miR-137 and OCT4 or SOX2 [47]. OCT4 and SOX2 are both pluripotent transcription factors with key roles in stem cell maintenance and differentiation [49,50,51]. Boyer et al. found that OCT4 and SOX2 co-occupy the promoter of mir-137 [51]. Direct binding to the 2.5 kb upstream region of miR-137 with MeCP2 and Sox2 in adult NSCs and H3-K4-Mono-Me and H3-K9-Tri-Me epigenetic modification induced by MeCP2 caused downregulation of miR-137 expression targeting Ezh2, a histone methyltransferase, thereby increasing proliferation and decreasing neuronal differentiation [46]. However, Channakkar et al. found that miR-137 overexpression in hNSCs increases OCT4 and SOX2 expression [47]. Therefore, miR-137 and OCT4 or SOX2 have been suggested to form a feed-forward self-regulating loop involving OCT4 and SOX2 as key components [51,52], through which miR-137 correctly regulates neurogenesis.

2.5. miR-153

miR-153 is highly conserved in zebrafish and humans and is highly expressed in the hippocampus in young rats and infants; it additionally plays a major role in adult neurogenesis [53,54]. According to a study by Qiao et al., endogenous miR-153 expression is downregulated with decreasing differentiation of neurons and is reduced with increasing numbers of NSC passages in vitro [55] However, the number of developed neurons has been found to increase after forced miR-153 expression in late NSCs, thus suggesting that miR-153 promotes neuronal differentiation.
Additionally, miR-153 has been found to control the Notch signaling pathway by preventing the translation of Jagged1 and Hey2. In vivo results demonstrated that overexpression of miR-153 in the hippocampus of old mice increased neurogenesis and cognitive impairment [55], implying the potential target for treatment.

2.6. miR-335-3p

miR-335-3p, whose gene is located on chromosome 7q32.2, is crucial in the emergence of a variety of cancers and tumors. This miRNA functions as a critical regulator of cell fate [56]. Furthermore, miR-335-3p is involved in a multitude of malignant tumors of the central nervous system (CNS), including well-known diseases such as astrocytoma and glioma [57]. Additionally, miR-335-3p specifically targets FMRP Translational Regulator 1 (Fmr1) [58], which synchronizes the migration of cortical neurons with the growth of the brain in developing mice [59]. In addition, evidence indicates that knockout of Fmr1 in adult NSCs decreases the expression of p53 [60], whereas the p53 signaling pathway promotes the self-renewal of NSCs [61].
As an important transcriptional factor, FoxM1 markedly influences cell proliferation and consequently controls neuroblastoma cell differentiation [62]. Jia et al. found that FoxM1 regulates miR-335-3p expression, as gain or loss of FoxM1 expression respectively increased or decreased miR-335-3p levels [63]. The expression of miR-335-3p also increased after restoration of FoxM1 expression in cells with FoxM1 knockdown. Data suggest that FoxM1 positively controls miR-335-3p expression. Notably, stimulation of the p53 signaling pathway via FoxM1-mediated stimulation of miR-335-3p may inhibit NSC differentiation and promote NSC self-renewal. The expression of FoxM1 in NSCs substantially decreases with increasing differentiation, as demonstrated by qPCR and Western blot analysis; these findings are compatible with the observed decline in the expression of miR-335-3p. According to the aforementioned findings, miR-335-3p is initially activated by FoxM1 and decreases the expression of Fmr1 before activating the p53 signaling pathway, thus increasing NSC self-renewal and preventing differentiation.

2.7. miR-410-3p

Accumulating evidence from animal studies has demonstrated that anesthetic exposure affects long term neurocognitive function in the developing brain [64,65,66]. Early research indicated that exposure to sevoflurane in the second trimester of pregnancy inhibits NSC development and triggers early death [67,68,69]. This mechanism may involve miR-410-3p, which is easily influenced by sevoflurane.
An uncommon neurodegenerative condition termed dentatorubral and pallidoluysian atrophy can be caused by mutations in the ATN1 gene. Atrophin-1 (ATN1) is extensively expressed in brain tissues [70] and recently has been suggested to be crucial for the maintenance of NSCs [71]. According to a report by Zhang, frequent exposure to sevoflurane may upregulate the expression of miR-410-3p and downregulate the expression of ATN1 [72]. Double luciferase reporter assays have revealed that ATN1 is the direct target of miR-410-3p.
Primary hippocampal NSCs harvested from the middle period in pregnant rats have been used to study the effects of sevoflurane on the differentiation of hippocampal NSCs. Repeated sevoflurane exposure has been found to give rise to the early differentiation of NSCs and the proliferation of astrocytes in the brains of developing offspring, thus resulting in a long-term decrease in hippocampal neurons. Lentivirus-mediated overexpression of ATN1 may diminish the effects of sevoflurane, and miR-410-3p inhibition via lentivirus transfection might improve NSC differentiation and ATN1 expression, thus indicating that ATN1 mediates the effects of miR-410-3p on hippocampal neurogenesis during sevoflurane exposure.

2.8. miR-145

A novel role of miR-145 in the neuronal differentiation of NSCs, as reported by Morgado et al., is inhibition of the self-renewal and pluripotency of human ESCs through the Sox2-Lin28/let-7 signaling pathway network [73]. Sox2, the core transcription factor of ESCs, is necessary for the early differentiation of NSCs, and its knockout can disrupt the differentiation of neurons [74,75]. A delicate balance of Sox2 levels is essential for normal neuron development. A previous report from Cimadamore et al. showed that Sox2 promotes NSC proliferation through Lin28/let-7b miRNA [74], and Sox2 is essential for the maintenance of optimal levels of Lin28, a developmental regulatory RNA binding protein that selectively inhibits the expression of let-7b miRNA. Lin28 and let-7b substantially mediate the function of Sox2 in the proliferation of NSCs.
Morgado et al. reported that Sox2 and Lin28 are direct targets of miR-145, and Sox2, Lin28, and let-7b are all regulated by miR-145, thus forming a double negative feedback loop [73]. However, miR-145 directly downregulates Sox2, thus further decreasing the expression of Lin28. Moreover, miR-145 directly decreases Lin28 expression. The low level of Lin28 results in an increase in let-7b. The above results have indicated that, during neurogenesis, the miR-145/Sox2-Lin28/let-7 signaling pathway might affect the differentiation of NSCs.

2.9. Other miRNAs

Previous studies have demonstrated that NSCs transplanted into the SGZ of hippocampi after denervation surgery exhibit robust neuronal differentiation and proliferation [76], implying that the microenvironmental changes in the denervated hippocampus facilitated neurogenesis. The following research revealed ncRNA expression patterns and neurogenesis-associated miRNAs in local exosomes using RNA sequencing and bioinformatic analysis of exosomes. The expression of miR-6324 and miR-3559-3p is notably higher in brains from the ectoderm than in tissue from other organs and is particularly high in the hippocampus. Additionally, overexpression of miRNAs miR-6324 and miR-3559-3p in hippocampal NSCs promotes neuronal differentiation while suppressing proliferation [77].
Through online database analyses and RNA-seq, we have also found several miRNAs that affect the proliferation and differentiation of NSCs, including miR-103-3p, miR-130-3p, and miR-107-3p. Reports have indicated that miR-103-3p controls NSC proliferation, differentiation, and apoptosis. Moreover, Ndel1 has been confirmed to be a novel miR-103-3p target whose overexpression induces NSC differentiation into neurons via the Wnt/β-catenin pathway, thus increasing NSC proliferation and decreasing NSC death [78]. These findings suggest a role of miR-103-3p/Ndel1 in NSC biological processes. By controlling the AKT/PI3K pathway, Acsl4 is targeted by miR-130a-3p to mediate the neuronal differentiation of NSCs [79]. Similarly, forced expression of miR-107-3p suppresses the neuronal differentiation of NSCs [80]. These findings have enabled deeper understanding of neurogenesis and might aid in the development of novel therapies for CNS diseases.
Table 1. Effects and mechanisms of various microRNAs in NSCs.
Table 1. Effects and mechanisms of various microRNAs in NSCs.
NameBiological FunctionTargetReferences
miR-124Inhibit apoptosis of NSCs/NPCs
Promote survival of hippocampal granule neurons and the normal development of axons
BCL2L13
Lhx2
[81]
miR-132Inhibit neuronal differentiation of ESs
Inhibit proliferation of radial glial cells and promote oligodendrocytes to differentiate
MECP2
P250GAP
[21,82]
miR-145Inhibit NSCs from differentiatingSOX2[73]
let-7fPromote NSCs to differentiate into astrocytes and neuronsLin28
Colin
[83]
miR-410-3pEarly differentiation of developing offspring brain NSCsATN1[72]
miR-335-3pMaintain self-renewal and inhibit NSCs from differentiatingFmr1[63]
miR-140-5pInhibit NSCs from differentiating into neurons
Promote differentiation into astrocytes
PROX1[34]
miR-137Promote neuronal differentiation of hiNSCsMEF2A
OCT4
[47]
miR-9Inhibit proliferation of NSCs and promote neuronal differentiation of adult NSCsTLX[84]
let-7bInhibit proliferation of NSCs and promote differentiation of adult NSCsTLX
cyclinD1
[85]
miR-137Promote proliferation of NSCs and inhibit neuronal differentiation of adult NSCs
Inhibit dendritic growth and the formation of dendritic spines in adult hippocampal newborn neurons
Ezh2
Mib-1
[52]
miR-184Drive symmetrical division of aNSCs and promote the self-renewal and proliferation of aNSCsNumb1[86]
miR-25Promote the proliferation of aNSCsFoxO3[87]
miR-17-92
cluster
Promote proliferation of adult NSCs and promote newborn neurons to differentiateSgk1[88]
miR-195Promote proliferation of adult NSCs and inhibit neurons from differentiatingMBD1[89]
miR-19Promote migration of newborn neurons in the adult hippocampusRapGEF2[90]
miR-I5aInhibit dendritic maturation of newborn neurons in the adult hippocampusBDNF[91]
miR-153Promote neurogenesis and cognitive abilityHey2, Jagged1[55]
miR-199 and miR-214Promote neuronal differentiation and induce neuronal progenitor proliferation and survival in prenatal developmentPAK4, PTEN[31]

3. Regulation of lncRNAs and the Differentiation of Hippocampal NSCs

lncRNAs are defined as a class of RNA molecules that do not encode proteins and whose transcripts are longer than 200 nucleotides. lncRNAs usually have the following characteristics: (1) lncRNA sequences are usually long and have an mRNA-like structure. (2) lncRNAs are not highly conserved, but the local sequences, and consequently their secondary and tertiary structures, are conserved; thus. lncRNAs are functionally conserved. (3) Expression of lncRNAs is relatively low. (4) Most lncRNAs show clear spatial–temporal expression specificity during differentiation and development. (5) Unlike miRNAs, lncRNAs do not have a universal mode of interaction. They can bind RNAs, DNAs, and proteins, and this binding can be enhanced or inhibited [92].
lncRNAs influence gene expression at the epigenetic, transcriptional, and post-transcriptional levels, and consequently are involved in various cellular processes, including cell division, survival, apoptosis, and motility [93]. Their localization enables diverse functions. For example, lncRNAs can function as miRNA “sponges” in the cytoplasm by binding miRNAs and releasing them from their 3′UTR binding sites; consequently, mRNA stability is modulated, and miRNA activity and translation in the nucleus are inhibited. [94,95]. Nuclear lncRNAs regulate chromatin and gene transcription [96]. The regulation of lncRNAs by transcriptional variable splicing is currently unclear, and its relationship to the regulation of neural differentiation requires further study. Table 2 summarized the lncRNAs reported to be related with neurogenesis.

3.1. Rik-201 and Rik-203

The lncRNA C130071C03Rik (mouse homologue of LINC00461) contains five transcripts (splice variants), among which the lncRNAs Riken (Rik)-201 and Rik-203 control neural differentiation. The lncRNA C130071C03Rik has been reported to be specifically expressed in the ependymal region of the spinal cord at embryonic days 11.5 and 13.5 in mice, and has been found to be expressed at higher levels in neural tissues than other tissues [97]. In comparison to other tissues, the hippocampus has an elevated level of Rik-203, which increases throughout neural development [98]. Recently, Zhang et al. found that Rik-203 and Rik-201 are present at high levels in the brain and promote mESC differentiation into neurons [99]. In addition, through RNA pull-down experiments, miR-467a-3p and miR-96 have been found to bind Rik-203 and Rik-201, respectively, thus indicating that Rik-203 and Rik-201 may serve as competing endogenous RNAs that inactivate miR-467a-3p and miR-96, respectively. Analysis using online miRNA target predication tools has indicated that Sox6 is a target of both miR-467a-3p and miR-96, and luciferase reporter assays have indicated that miR-467a-3p and miR-96 bind and downregulate Sox6 expression. Sox6 specifically regulates dopamine neurons during neural development and plays a regulatory role in the early differentiation of hippocampal NSCs [100,101]. Furthermore, gain or loss of function experiments have shown that Rik-203 and Rik-201 knockdown and overexpression of miR-96 and miR-467a-3p both inhibit neural differentiation. Sox6 reverses the suppression of differentiation in NSCs by upregulating miR-96 and miR-467a-3p and downregulating Rik-203 and Rik-201. These results indicate that Rik-203 and Rik-201 act as competing endogenous RNAs in the functional inhibition of miR-467a-3p and miR-96, respectively, and modulate the expression of Sox6, thereby further regulating NSC differentiation [99].
Sevoflurane has been extensively used in clinical anesthesia, and early childhood exposure to sevoflurane has been associated with neural abnormalities in both humans and animals according to prior research [98,102]. Sevoflurane exposure potently inhibits NSC self-renewal and differentiation in vitro [103], thus leading to neuronal loss and cognitive impairment in young animals. Sevoflurane has also been found to decrease Rik-203 levels in mouse hippocampal tissue and neural progenitor cells. Suppression of Rik-203 levels in neural progenitors decreases not only Sox1 and Nestin expression, but also the number of Sox1 positive cells. miR-101a-3p strongly binds Rik-203, as evidenced by RNA-RNA pull-down experiments. Sevoflurane, Rik-203 suppression, and miR-101a-3p overexpression all decrease GSK-3β levels; consequently, a cascade of miR-101a-3p and GSK-3β may be involved in the Rik-203-mediated regulation of neural development [98]. Rik-203 also attenuates neuronal differentiation via inhibition of downstream miR-466l-3p. Sevoflurane decreases Rik-203 levels, thereby resulting in the release of miR-466L-3p from Rik-203. When miR-466L-3p is released, it specifically targets and decreases the concentration of BDNF and subsequently inhibits neuronal development [104]. The results above suggest that Rik-203 is a potential target that may be important in the prevention of anesthesia-induced neurotoxicity.

3.2. Peg13

The lncRNA Peg13 was first reported to be differentially expressed in cerebral vascular endothelial cells in an ischemic glucose deficiency model in 2016, thus suggesting its potential role in nervous system injury [105]. Recently, Peg13 has been shown to regulate social and sexual interactions in mice, and mice lacking Peg13 exhibit same-sex attraction-like behavior. These mice also exhibit a deficiency in pup retrieval behavior, elevated anxiety, and diminished activity and curiosity [106]. Peg13 has been found to be involved in controlling mouse mating choices and modulating sevoflurane-associated neurotoxicity against NSCs [107,108]. Reports have indicated that an axis between miRNAs and target genes mediating Peg13 sponges is involved in the regulation of biological function. Exposure to sevoflurane significantly downregulates the expression of Peg13 and Sox13 and upregulates the expression of miR-128-3p in neural stem cells. Peg13 serves as a molecular sponge for miR-128-3p, thereby maintaining the expression of Sox13 in NSCs and decreasing sevoflurane-associated neurotoxicity [103]. Peg13 bound to miR-490-3p upregulates Psmd11, thus inactivating the Wnt/β-catenin pathway and alleviating the progression of epilepsy [107]. Mechanistically, Peg13 may function as a sponge for miR-20a-5p, thereby increasing the expression of XIAP and lessening serious brain injury caused by hypoxia/ischemia in newborn mice [109]. The Peg13/miRNA/target gene axis may have promising applications as a potential therapeutic target in neurological diseases.

3.3. lncRNA1230

lncRNA1230 is a long intergenic non-coding RNA. Large-scale microarray data from Guttman et al. [107] indicate that the expression of lncRNA1230 is associated with differentiation of the neuroectoderm in mouse embryonic stem cells, thus suggesting its involvement in the control of neural lineage determination. Forced ectopic expression of lncRNA1230 significantly attenuates the ability of mouse embryonic stems to form neural cells, whereas knockdown of lncRNA1230 promotes the conversion of mouse embryonic stem cells toward certain NPCs. The lncRNA1230-induced mechanism of inhibition in the transformation of mouse ESCs to NPCs involves decreasing trimethylation of histone 3 lysine 4, thus inhibiting the binding of WD repeat domain 5 (WDR5) to the promoter regions of neurogenesis-related genes [108]. Notably, lncRNA1230 may play a crucial role in the neural fate of stem cells.

3.4. Pnky

Pnky is a highly conserved neuro-specific lncRNA localized primarily in the nucleus. In 2015, Ramos et al. [110] found that Pnky knockdown accelerates the differentiation of NSCs into mature neurons and significantly decreases the numbers of NSCs. Therefore, Pnky plays a critical regulatory role in NSC differentiation. Reports have indicated that Pnky inhibits the differentiation of NSCs by binding and interacting with polypyrimidine bundle binding protein (polypyrimidine tract-binding protein, PTBP1) and consequently inhibiting the splicing and expression patterns of key mRNAs in neuronal stem cells [111]. Pnky has also been shown to be a key regulator of NSC migration through modulation of the splicing and export of target mRNAs [112].
Although lncRNAs play important roles in cell biology, few have been shown to regulate in vivo development, particularly cis and trans-regulation. In one example, after knockout of Pnky in developing cortical cells, the expression of POU Class 3 Homeobox 2 (Pou3f2) decreased and differentiation of NSCs was promoted, although the expression of Pou3f2 was not completely lost with the deletion of the Pnky gene. Pnky promotes the expression of Pou3f2 and regulates the differentiation of NSCs through trans-regulatory mechanisms [113].
In-depth knowledge of Pnky may potentially allow for its use in the treatment of neural disorders. Lin et al. [114] designed and synthesized an MRI visualization nano-drug that immobilizes a mixture of siRNAs targeting Pnky on micelle surfaces and promotes the directional differentiation of NSCs into neurons by downregulating the level of Pnky, thus repairing cerebral infarctions. These results indicate the great potential of nanomedicines targeting Pnky in NSC-based therapies, particularly for stroke.

3.5. Neat1

The overexpression and knockdown of the lncRNA Neat1 respectively promotes and inhibits both spinal cord NSC migration and differentiation into neurons [115]. Neat1 has been reported to be the key regulator of Wnt/β-catenin [116], the key pathway affecting the proliferation of NSCs [117]. The expression of Neat1 is regulated by miR-124, thus resulting in activation of the Wnt/β-catenin pathway during spinal cord injury regeneration [115]. Recent studies have indicated that the lncRNA Neat1 mediates the proliferation of neural stem cells via the Neat1-let 7 b-P21 axis [118]. Neat1 may also mediate several pathways regulating neurogenesis. Beyond promoting NSCs to differentiate into neurons, Neat1 promotes NSCs to differentiate into oligodendrocytes. In 2019, Katsel et al. found the diminished expression of Neat1 in the brains of patients with schizophrenia, with a greatly diminished number of oligodendrocytes also thus observed [119]. In terms of the above reports, there exists a feedback loop between lncRNA and miRNA, forming a mutual regulation relationship. As a competing endogenous RNA, miR-124 is the target of Neat1 and conversely controls its expression, and together they regulate downstream gene expression and signal activity.

3.6. GAS5

The lncRNA GAS5 has been extensively studied in tumors and is thought to potentially function as an anti-oncogene [120]. Although research on the importance of GAS5 in neurological disorders such as AD is scarce, GAS5 levels are significantly higher in patients with AD than controls [121]. The function of GAS5 in neurogenesis has also been investigated, with GAS5 being found to promote hippocampal NSC differentiation into neurons. In rats with cholinergic injury, overexpression of GAS5 enhances learning and memory in vivo [122]. GAS5 was found to be regulated by Lhx8, which has a specific effect on the development of the cholinergic nervous system and promotes cholinergic differentiation of hippocampal NSCs. Therefore, it has been hypothesized that an elevated GAS5 level in the hippocampus promotes NSCs in the hippocampal DG to differentiate into neurons.
Table 2. Effects and mechanisms of various lncRNAs on NSCs.
Table 2. Effects and mechanisms of various lncRNAs on NSCs.
NameBiological FunctionMechanismReferences
Sox2OTInhibit NSC proliferation and promote neuronal differentiationLink with Sox2, interact with YY1[123]
RMSTPromote the formation and development of nerve cellsTarget Sox2[124,125]
Kdm2bPromote the formation and development of nerve cellsCombine with hnRNPAB and activate Kdm2b expression[126]
PauparPromote the formation and development of nerve cellsCombine with local genes Pax6 and KAP1[127]
Gm21284Inhibit NSC proliferation while promoting NSC differentiationInteract with miR-30e-3p, miR-147, and miR-431[128]
1604Enhance neural differentiationmiR-200c/ZEB1/2 axis[129]
Rik-201Promote neural differentiationRegulated by C/EBPβ and target miR-96/Sox6[99]
Rik-203Promote neural differentiationRegulated by C/EBPβ and target miR-467a-3p/Sox6 and miR-101-3a/GSK-3β[98,99]
Malat1Enhance neural differentiationActivate ERK/MAPK, inhibit PPAR/p53[130]
PnkyInhibit neural differentiation and the formation and development of nerve cellsInteract with PTBP1[110]
IncR492Inhibit neural differentiationInteract with HuR and activate Wnt signaling[131]
BDNF-ASInhibit eNSC-derived neurite outgrowthTarget TrkB signaling pathway[132]
UCA1Promote NSC differentiation to astrocytesmiR-1/Hes1[133]
OPCBoost oligodendrogenesisRegulated by OLIG2[134]
IncOL1Boost oligodendrogenesisForm a complex with Suz12[135]
Inc158Boost oligodendrogenesisPromote NFIB expression[136]
Neat1Boost oligodendrogenesisActivate the Wnt/β-catenin pathways[115]
Pcdh17itFunction as an oligodendrogenesis markerUnknown[137]
OLMALIN/-ASModulate oligodendrocyte maturationMay affect several genes, Target multiple genes, such as HDAC9, SOX4, GPR126 and EGR1[138]
MAG3Inhibit neurogenesisAffect Notch or Wnt/β-catenin signaling pathway, miR-128-3p/ATRA/cAMP/CREB axis[139,140]
lncRNA1230Attenuate NSCs’ ability to form neural cellsInteract with Wdr5[141]
LINGO-1Promote neurogenesisDownregulate miR-15b-3p and promote Wnt5a expression[142]
Peg13Promote neurogenesisSponge microRNA-128-3p to preserve Sox13 expression[103]
GAS5Promote hippocampal NSC differentiation into neuronsUnknown[121,122]

4. Regulation of circRNAs in Hippocampal NSCs

circRNAs are a newly described RNA species formed by the reverse splicing of linear genes. These stable closed circular RNA molecules are widely conserved and expressed in eukaryotes [143]. circRNAs act as molecular “sponges” for miRNAs and regulate gene transcription, affect protein synthesis and function, and participate in the regulation of intercellular signaling pathways [144]. Clustering of circular RNA sequencing data from different regions of the mouse brain (olfactory bulb, prefrontal cortex, hippocampus, and cerebellum) has indicated that expression of these RNAs is brain region-specific [145], and circRNAs are highly correlated with CNS diseases [146,147].
lncRNAs, miRNAs, and circRNAs are inextricably linked; for example, circRNAs and lncRNAs can negatively regulate the expression of miRNA through an miRNA sponge mechanism, and can further interfere with the complementary base pairs of target mRNAs [148], thus modulating gene expression and interfering with the occurrence and development of disease [149]. Many ncRNAs do not play independent roles in cells but mutually interact and collaboratively influence biological functions. Table 3 summarized the circRNAs reported to be related with NSCs.

4.1. circHIPK2

Wang et al. showed that circHIPK2 plays a negative role in the neural differentiation of NSCs. In vitro silencing of circHIPK2 causes direct differentiation of NSCs to neurons but not to astrocytes. Transplantation of NSCs with downregulated circHIPK2 has been found to facilitate neural plasticity and functional recovery after stroke; therefore, circHIPK2 may be a potential target for stroke treatment [150].
RNA sequencing has indicated that circHIPK2 targets and increases the expression of Smox. Silencing any of them facilitates neuronal differentiation in vitro but does not affect differentiation into astrocytes [150]. Thus, the effect on neuronal differentiation after silencing of circHIPK2 is potentially mediated by inhibition of the expression of Smox, thereby further promoting nerve recovery after stroke. However, the identities of the intermediary molecules between circHIPK2 and Smox remain elusive.
miR-124 was proposed to have an intermediary role after Huang et al. discovered that circHIPK2 functions as an endogenous sponge for miRNA-124—a miRNA that regulates autophagy and endoplasmic reticulum stress [146]. Interestingly, miRNA-124 is involved in the regulation of neural differentiation [151,152], and Smox has been found to be the target gene of miRNA-124 in the development of H. pylori-associated gastric cancer [153]. The circHIPK2/miR-124/Smox axis is believed to play an important role in neurogenesis, although this possibility has not been confirmed experimentally.

4.2. circ-TTC3

circRNA tetratricopeptide repeat domain 3 (circ-TTC3) has been studied in a limited number of diseases but is known to be upregulated in hypoxic cardiomyocytes and to protect against myocardial infarction-induced cardiomyocyte apoptosis via the miR-15b-Arl2 regulatory cascade [154]. Additionally, circ-TTC3 binds miR-449a, thus activating the NFⱪ-B and PI3K/AKT pathways and ameliorating hypoxia-induced damage in HaCaT cells [155]. Moreover, during acute kidney injury caused by sepsis, circ-TTC3 alleviates inflammation and oxidative stress via the miR-148a/Rcan2 axis [156].
The effect of circ-TTC3 on NSC differentiation has been demonstrated by Yang et al. [157]: depletion of circ-TTC3 in NSCs increases proliferation and neuronal differentiation. circ-TTC3 appears to sponge miR-372-3p and then target and enhance Toll-like receptor 4 (TLR4) expression in NSCs. Evidence also indicates that TLR4 is involved in promoting NSC differentiation into astrocytes and neurons during stroke progression [158]. Depletion of circTTC3 was found to significantly decrease neurological scores, brain water content, and cerebral infarction in a middle cerebral artery occlusion/repression stroke model. Moreover, the effect of circTTC3 on stroke and NSC biology has been confirmed to be mediated through the miR-372-3p/TLR4 axis.

4.3. Acbd6

The circRNA Acbd6 is spliced from exons derived from Acbd6 on the positive strand of chromosome 13 (from 73, 239, 821–73, 265, and 813 bp). Virtually no reports have investigated circAcbd6, and few studies have examined the host genes of circAcbd6. However, members of the ACBD family have been reported to play important roles in regulating viral replication, organelle organization, self-renewal of stem cells, and protein acylation [159]. A modular protein in mammalian cells, acyl-coenzyme A binding domain-containing member 6 (Acbd6) is found in the spleen, placenta, and embryonic-like stem cells produced from the placenta, bone marrow, cord blood, and circulating CD34+ progenitors [159,160]. In a recent study, a particular region of the SGZ that harbors a subpopulation of mature NSCs showed a substantial expression of circAcbd6 in neural tissues. Moreover, the finding from our lab shows that the forced expression of circAcbd6 promotes hippocampal NSC differentiation into neurons and even cholinergic neurons [161]. In vivo results suggest that circAcbd6 may greatly enhance learning and memory ability, indicating its potential value in the treatment of AD.
Table 3. Effects and mechanisms of various circRNAs in NSCs.
Table 3. Effects and mechanisms of various circRNAs in NSCs.
NameBiologic FunctionMechanismReferences
circHIPK2Inhibit proliferation and differentiation of NSCsCombine miR-124 and regulate Smox expression[146]
circRNA TTC3Inhibit proliferation and differentiation of NSCsRegulate miR-372-3p/TLR4 axis[157]
hsa-circ-0002468Regulate neuronal differentiationModulate miR-561/E2F8 axis[162]
circRNA Acbd6Promote NSC differentiation into cholinergic neuronsModulate miR-320-5p-Osbpl2 axis[161]
circ-0005835Inhibit NSC proliferation and differentiation to neuronsSponge miR-576-3p[163]

5. Studies of ncRNAs in Hippocampus-Related Diseases

5.1. Alzheimer’s Disease

Alzheimer’s disease (AD), a progressive neurodegenerative disease involving multiple damaged brain regions (including the hippocampus), is the main cause of dementia in older people globally. It is characterized by the deposition of Aβ plaques, neurofibrillary tangles, and chronic inflammation, thus leading to a loss of neurons; this loss is closely associated with the accumulation of Aβ and the hyperphosphorylation of Tau [164,165,166]. Many studies in patients with AD have reported many upregulated miRNAs in the brain (including miRNA-7, miRNA-9, miRNA-34a, miRNA-125b, miRNA-146a, and miRNA-155) [167,168]. According to a recent study in Nature Neuroscience, a substantial correlation exists between AD characteristics and circHOMER1, circMAN2A1, and circKCNN2. circHOMER1, circKCNN2, and other circRNAs were discovered in all four meta-analyses; two quantitative AD characteristics have also been associated with these eight circRNAs (circDNAJC6, circFMN1, circDOCK1, circERBIN, circHOMER1, circKCNN2, circMAN2A1, and circST18) [169,170]. Idda et al. analyzed and reported the abnormal expression of lncRNA in AD and found that BACE1-AS, 51A, and NAT-Rad18 are important ncRNAs that are involved in the onset of AD [171].
In view of these abnormally expressed ncRNAs and important target genes, anti-miRNA strategies, anti-NF-kB strategies, and other strategies for the treatment of AD have been proposed. RNA oligonucleotides, which have recently been shown to potently suppress miRNAs, are being developed as anti-miRNA drugs, and techniques using viral vector delivery systems have been found to maintain miRNA-associated homeostasis in brain tissue [172]. The ncRNAs involved in AD are summarized in Table 4.

5.2. Epilepsy

Epilepsy, a common chronic disease affecting the nervous system, is characterized by recurrent spontaneous seizures and affects approximately 1% of the global population. The hippocampus is an important area in relation to the mediation of epileptic discharges [217]. Although great progress has been made in the development of antiepileptic drugs, some patients with epilepsy are not sensitive to drug treatment, and many patients have adverse reactions to antiepileptic drugs. Epilepsy is attributed to an imbalance between inhibitory and excitatory neurons in the CNS that leads to the “abnormal discharge” of neurons [218].
In many cases, epilepsy occurs alongside a neurodegenerative disorder, such as AD. Some patients with epilepsy experience memory impairment, whereas 16% of patients with AD develop epilepsy during severe dementia [219]. Post-stroke epilepsy can occur shortly after an ischemic attack or later, and acute ischemia can increase the concentrations of extracellular glutamate, thereby leading to periodic epileptiform discharge in neurons [220,221,222].
More and more ncRNAs have been found to be closely involved in the genesis or pathophysiological processes of epilepsy, exhibiting the value of biomarkers for diagnosis and prognosis [223,224]. Taking miRNA as an example, accumulating evidence indicates that the miRNA landscape is dysregulated in epilepsy, igniting prospects for the treatment of epilepsy [225,226]. The strategy aiming at miRNA involves administration of chemically engineered antisense oligonucleotides (ASOs, mimics or antagomirs) targeting the given miRNAs, which would thus be reproduced or inhibited. For example, miRNA-124 is a very small important RNA molecule whose expression is highest in the hippocampus. During the onset of epilepsy, by the addition of an activator, levels of miRNA-124 may increase, thus regulating CREB1 and other genes and consequently decreasing the degree of seizures and preventing excessive neuronal apoptosis [227,228]. Jimenez-Mateoset et al. found that in vivo administration of antagomirs targeting miR-134 could upregulate spine volume in the hippocampus and attenuate the degree of seizures in epileptic mice [229]. These studies provide evidence to support the use of specific ASOs targeting miRNAs or other ncRNAs for anticonvulsant effects.

5.3. Stroke and I/R Injury

The hippocampus is usually the victim in the process of stroke and ischemia reperfusion (I/R) injuries, which are often followed by episodic memory impairments. Increasing evidence indicates that ncRNAs exert crucial roles in the pathophysiological processes of stroke or I/R injury, including neuroinflammation, oxidative stress, excitotoxicity, and apoptosis, as summarized in Table 5. With the increasing incidence of stroke and I/R injury, studies examining the related pathogenesis have found that regulation of ncRNAs promotes the reversal of these diseases. For instance, the lncRNA MEG3 inhibits neuronal death and neurological problems in mice with ischemic stroke through the MEG3/miR-378/Grb2 axis and inhibits neuronal autophagy and neurodegeneration by silencing Grb2. In vitro studies have revealed that MEG3 specifically binds miR-378 and increases Grb2 expression, thus inhibiting the activation of the Akt/mTOR pathway and protecting neurons against autophagy and neurological damage [230]. Studies on animals showed that using tailored exosomes enriched with miR-17-92 clusters led to improvement of neurological function and enhancements of neurogenesis during brain stroke via activation of the PI3K/AKT/mTOR signaling pathway [231]. Intravenous administration of an antagomir against miR-15a/16-1 significantly alleviated ischemic damage via upregulation of antiapoptotic proteins and inhibition of proinflammatory molecules [232]. lncRNAs and circRNAs were also reported to play critical roles in the progression of ischemic stroke, thus likewise acting as potential therapeutic targets for stroke. Therefore, treatments which target ncRNAs or make use of ncRNA molecules might help to improve the prognosis of stroke and I/R injury.

6. Limitations and Prospects

Current research has utilized the characteristics and advantages of ncRNAs (especially miRNAs and circRNAs) as new biological markers to screen for several diseases. These ncRNAs have major roles in disease initiation and progression and great value in disease diagnosis, treatment, and prognostication. Interfering with the regulation of ncRNAs can help control disease progression; therefore, regulation of the expression of key ncRNAs may provide hope for disease prevention and treatment and the identification of targets for new drug development. At present, the future of ncRNA research is promising. Currently, in the treatment of neurological illnesses, ncRNAs can be regulated through a range of physical techniques, such as cranial magnetic resonance and electroacupuncture. However, recent studies have shown that the development of NSCs is a precisely controlled process whose influencing factors are not regulated by single signaling pathways but are instead closely associated with a variety of different cellular signaling pathways and cellular electrophysiological characteristics. The regulatory network is complex, involving chromatin remodeling, histone modification, DNA methylation, X chromosome inactivation, and regulation of ncRNA at the translational level. It is also associated with changes in the extracellular microenvironment. Recent studies have shown that ncRNAs play important roles in the differentiation of NSCs. However, several limitations must be noted. Research on the effects of ncRNAs on NSC differentiation has been relatively simple and limited to examining the effect of a certain type of ncRNA on NSC differentiation. The specific mechanisms remain to be fully elucidated, and some influencing factors remain hypothetical. In addition, few reports have described the formation of common regulatory networks among ncRNAs, mRNAs, and the microenvironment affecting the differentiation of NSCs. Furthermore, the effects of ncRNAs on the differentiation of NSCs involve multiple molecules. Although in recent years lncRNAs have been confirmed to act as miRNA sponges, few studies have assessed the effects of ncRNAs on the nervous system. Moreover, the construction of lncRNA-miRNA-mRNA regulatory networks is imperfect, and further research in this field is urgently needed. Nonetheless, ncRNAs are important regulators affecting the differentiation of NSCs and may enable the development of potentially valuable novel treatments for nervous system diseases.

Author Contributions

Conceptualization, X.Z.; methodology and validation, Z.T. and W.L.; formal analysis, Z.T. and X.C.; writing—original draft preparation, Z.T. and W.L.; writing—review and editing, X.Z. and Q.Z. All authors have read and agreed to the published version of the manuscript.

Funding

The present study was supported by the National Natural Science Foundation of China (grant no. 31171038), Jiangsu Natural Science Foundation (grant no. BK2011385), Jiangsu ‘333’ program funding (grant no. BRA2016450), the Application Research Project of Nantong City (grant no. MS12017015-3), and the Training Program of Innovation and Entrepreneurship for Graduates (No. 265, 202210304043Z).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

No data support in publicly archived datasets.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, J.H.; Liu, S.; Zhou, H.; Qu, L.H.; Yang, J.H. starBase v2.0: Decoding miRNA-ceRNA, miRNA-ncRNA and protein-RNA interaction networks from large-scale CLIP-Seq data. Nucleic Acids Res. 2014, 42, D92–D97. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Weng, M.; Wu, D.; Yang, C.; Peng, H.; Wang, G.; Wang, T.; Li, X. Noncoding RNAs in the development, diagnosis, and prognosis of colorectal cancer. Transl. Res. 2017, 181, 108–120. [Google Scholar] [CrossRef] [PubMed]
  3. Matsui, M.; Corey, D.R. Non-coding RNAs as drug targets. Nat. Rev. Drug Discov. 2017, 16, 167–179. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Wang, J.; Zhu, S.; Meng, N.; He, Y.; Lu, R.; Yan, G.R. ncRNA-Encoded Peptides or Proteins and Cancer. Mol. Ther. J. Am. Soc. Gene Ther. 2019, 27, 1718–1725. [Google Scholar] [CrossRef] [PubMed]
  5. Zhang, P.; Wu, W.; Chen, Q.; Chen, M. Non-Coding RNAs and their Integrated Networks. J. Integr. Bioinform. 2019, 16. [Google Scholar] [CrossRef] [PubMed]
  6. Takouda, J.; Katada, S.; Nakashima, K. Emerging mechanisms underlying astrogenesis in the developing mammalian brain. Proc. Jpn. Acad. Ser. B Phys. Biol. Sci. 2017, 93, 386–398. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  7. Bak, M.; Silahtaroglu, A.; Møller, M.; Christensen, M.; Rath, M.F.; Skryabin, B.; Tommerup, N.; Kauppinen, S. MicroRNA expression in the adult mouse central nervous system. RNA 2008, 14, 432–444. [Google Scholar] [CrossRef] [Green Version]
  8. Smirnova, L.; Gräfe, A.; Seiler, A.; Schumacher, S.; Nitsch, R.; Wulczyn, F.G. Regulation of miRNA expression during neural cell specification. Eur. J. Neurosci. 2005, 21, 1469–1477. [Google Scholar] [CrossRef]
  9. Kapsimali, M.; Kloosterman, W.P.; de Bruijn, E.; Rosa, F.; Plasterk, R.H.; Wilson, S.W. MicroRNAs show a wide diversity of expression profiles in the developing and mature central nervous system. Genome Biol. 2007, 8, R173. [Google Scholar] [CrossRef] [Green Version]
  10. Lattanzi, A.; Gentner, B.; Corno, D.; Di Tomaso, T.; Mestdagh, P.; Speleman, F.; Naldini, L.; Gritti, A. Dynamic Activity of miR-125b and miR-93 during Murine Neural Stem Cell Differentiation In Vitro and in the Subventricular Zone Neurogenic Niche. PLoS ONE 2013, 8, e67411. [Google Scholar] [CrossRef]
  11. Boldrini, M.; Fulmore, C.A.; Tartt, A.N.; Simeon, L.R.; Pavlova, I.; Poposka, V.; Rosoklija, G.B.; Stankov, A.; Arango, V.; Dwork, A.J.; et al. Human Hippocampal Neurogenesis Persists throughout Aging. Cell Stem Cell 2018, 22, 589–599.e585. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Bergmann, O.; Spalding, K.L.; Frisén, J. Adult Neurogenesis in Humans. Cold Spring Harb. Perspect. Biol. 2015, 7, a018994. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Obernier, K.; Alvarez-Buylla, A. Neural stem cells: Origin, heterogeneity and regulation in the adult mammalian brain. Development 2019, 146, dev156059. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Vieira, M.S.; Santos, A.K.; Vasconcellos, R.; Goulart, V.A.M.; Parreira, R.C.; Kihara, A.H.; Ulrich, H.; Resende, R.R. Neural stem cell differentiation into mature neurons: Mechanisms of regulation and biotechnological applications. Biotechnol. Adv. 2018, 36, 1946–1970. [Google Scholar] [CrossRef] [PubMed]
  15. Faigle, R.; Song, H. Signaling mechanisms regulating adult neural stem cells and neurogenesis. Biochim. Biophys Acta 2013, 1830, 2435–2448. [Google Scholar] [CrossRef] [Green Version]
  16. Lagos-Quintana, M.; Rauhut, R.; Lendeckel, W.; Tuschl, T. Identification of novel genes coding for small expressed RNAs. Science 2001, 294, 853–858. [Google Scholar] [CrossRef] [Green Version]
  17. Calin, G.A.; Sevignani, C.; Dumitru, C.D.; Hyslop, T.; Noch, E.; Yendamuri, S.; Shimizu, M.; Rattan, S.; Bullrich, F.; Negrini, M.; et al. Human microRNA genes are frequently located at fragile sites and genomic regions involved in cancers. Proc. Natl. Acad. Sci. USA 2004, 101, 2999–3004. [Google Scholar] [CrossRef] [Green Version]
  18. Chen, P.Y.; Manninga, H.; Slanchev, K.; Chien, M.; Russo, J.J.; Ju, J.; Sheridan, R.; John, B.; Marks, D.S.; Gaidatzis, D.; et al. The developmental miRNA profiles of zebrafish as determined by small RNA cloning. Genes Dev. 2005, 19, 1288–1293. [Google Scholar] [CrossRef] [Green Version]
  19. Coutinho, L.L.; Matukumalli, L.K.; Sonstegard, T.S.; Van Tassell, C.P.; Gasbarre, L.C.; Capuco, A.V.; Smith, T.P. Discovery and profiling of bovine microRNAs from immune-related and embryonic tissues. Physiol. Genom. 2007, 29, 35–43. [Google Scholar] [CrossRef]
  20. Yoshimura, A.; Numakawa, T.; Odaka, H.; Adachi, N.; Tamai, Y.; Kunugi, H. Negative regulation of microRNA-132 in expression of synaptic proteins in neuronal differentiation of embryonic neural stem cells. Neurochem. Int. 2016, 97, 26–33. [Google Scholar] [CrossRef]
  21. Chen, D.; Hu, S.; Wu, Z.; Liu, J.; Li, S. The Role of MiR-132 in Regulating Neural Stem Cell Proliferation, Differentiation and Neuronal Maturation. Cell Physiol. Biochem. 2018, 47, 2319–2330. [Google Scholar] [CrossRef] [PubMed]
  22. Luikart, B.W.; Bensen, A.L.; Washburn, E.K.; Perederiy, J.V.; Su, K.G.; Li, Y.; Kernie, S.G.; Parada, L.F.; Westbrook, G.L. miR-132 mediates the integration of newborn neurons into the adult dentate gyrus. PLoS ONE 2011, 6, e19077. [Google Scholar] [CrossRef] [PubMed]
  23. Walgrave, H.; Balusu, S.; Snoeck, S.; Vanden Eynden, E.; Craessaerts, K.; Thrupp, N.; Wolfs, L.; Horre, K.; Fourne, Y.; Ronisz, A.; et al. Restoring miR-132 expression rescues adult hippocampal neurogenesis and memory deficits in Alzheimer’s disease. Cell Stem Cell 2021, 28, 1805–1821.e8. [Google Scholar] [CrossRef]
  24. Zhou, Z.D.; Kumari, U.; Xiao, Z.C.; Tan, E.K. Notch as a molecular switch in neural stem cells. IUBMB Life 2010, 62, 618–623. [Google Scholar] [CrossRef] [PubMed]
  25. Salta, E.; Lau, P.; Sala Frigerio, C.; Coolen, M.; Bally-Cuif, L.; De Strooper, B. A self-organizing miR-132/Ctbp2 circuit regulates bimodal notch signals and glial progenitor fate choice during spinal cord maturation. Dev. Cell 2014, 30, 423–436. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  26. Tillotson, R.; Bird, A. The Molecular Basis of MeCP2 Function in the Brain. J. Mol. Biol. 2019, 432, 1602–1623. [Google Scholar] [CrossRef]
  27. Zocher, S.; Overall, R.W.; Lesche, M.; Dahl, A.; Kempermann, G. Environmental enrichment preserves a young DNA methylation landscape in the aged mouse hippocampus. Nat. Commun. 2021, 12, 3892. [Google Scholar] [CrossRef]
  28. Pejhan, S.; Del Bigio, M.R.; Rastegar, M. The MeCP2E1/E2-BDNF-miR132 Homeostasis Regulatory Network Is Region-Dependent in the Human Brain and Is Impaired in Rett Syndrome Patients. Front. Cell Dev. Biol. 2020, 8, 763. [Google Scholar] [CrossRef]
  29. Klein, M.E.; Lioy, D.T.; Ma, L.; Impey, S.; Mandel, G.; Goodman, R.H. Homeostatic regulation of MeCP2 expression by a CREB-induced microRNA. Nat. Neurosci. 2007, 10, 1513–1514. [Google Scholar] [CrossRef]
  30. Zhang, X.M.; Huang, G.W.; Tian, Z.H.; Ren, D.L.; Wilson, J.X. Folate stimulates ERK1/2 phosphorylation and cell proliferation in fetal neural stem cells. Nutr. Neurosci. 2009, 12, 226–232. [Google Scholar] [CrossRef]
  31. Mellios, N.; Feldman, D.A.; Sheridan, S.D.; Ip, J.P.K.; Kwok, S.; Amoah, S.K.; Rosen, B.; Rodriguez, B.A.; Crawford, B.; Swaminathan, R.; et al. MeCP2-regulated miRNAs control early human neurogenesis through differential effects on ERK and AKT signaling. Mol. Psychiatry 2018, 23, 1051–1065. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Liang, H.B.; He, J.R.; Tu, X.Q.; Ding, K.Q.; Yang, G.Y.; Zhang, Y.; Zeng, L.L. MicroRNA-140-5p: A novel circulating biomarker for early warning of late-onset post-stroke depression. J. Psychiatr. Res. 2019, 115, 129–141. [Google Scholar] [CrossRef] [PubMed]
  33. Liang, H.B.; Lai, Z.H.; Tu, X.Q.; Ding, K.Q.; He, J.R.; Yang, G.Y.; Sheng, H.; Zeng, L.L. MicroRNA-140-5p exacerbates vascular cognitive impairment by inhibiting neurogenesis in the adult mouse hippocampus after global cerebral ischemia. Brain Res. Bull. 2022, 183, 73–83. [Google Scholar] [CrossRef] [PubMed]
  34. Ding, K.; Lai, Z.; Yang, G.; Zeng, L. MiR-140-5p targets Prox1 to regulate the proliferation and differentiation of neural stem cells through the ERK/MAPK signaling pathway. Ann. Transl. Med. 2021, 9, 671. [Google Scholar] [CrossRef] [PubMed]
  35. Oliver, G.; Sosa-Pineda, B.; Geisendorf, S.; Spana, E.P.; Doe, C.Q.; Gruss, P. Prox 1, a prospero-related homeobox gene expressed during mouse development. Mech. Dev. 1993, 44, 3–16. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Galeeva, A.; Treuter, E.; Tomarev, S.; Pelto-Huikko, M. A prospero-related homeobox gene Prox-1 is expressed during postnatal brain development as well as in the adult rodent brain. Neuroscience 2007, 146, 604–616. [Google Scholar] [CrossRef]
  37. Lavado, A.; Oliver, G. Prox1 expression patterns in the developing and adult murine brain. Dev. Dyn. Off. Publ. Am. Assoc. Anat. 2007, 236, 518–524. [Google Scholar] [CrossRef]
  38. Karalay, O.; Doberauer, K.; Vadodaria, K.C.; Knobloch, M.; Berti, L.; Miquelajauregui, A.; Schwark, M.; Jagasia, R.; Taketo, M.M.; Tarabykin, V.; et al. Prospero-related homeobox 1 gene (Prox1) is regulated by canonical Wnt signaling and has a stage-specific role in adult hippocampal neurogenesis. Proc. Natl. Acad. Sci. USA 2011, 108, 5807–5812. [Google Scholar] [CrossRef] [Green Version]
  39. Chen, J.; Liang, L.; Li, Y.; Zhang, Y.; Zhang, M.; Yang, T.; Meng, F.; Lai, X.; Li, C.; He, J.; et al. Naloxone regulates the differentiation of neural stem cells via a receptor-independent pathway. FASEB J. 2020, 34, 5917–5930. [Google Scholar] [CrossRef] [Green Version]
  40. Karalay, O.; Jessberger, S. Translating niche-derived signals into neurogenesis: The function of Prox1 in the adult hippocampus. Cell Cycle 2011, 10, 2239–2240. [Google Scholar] [CrossRef]
  41. Lavado, A.; Lagutin, O.V.; Chow, L.M.; Baker, S.J.; Oliver, G. Prox1 is required for granule cell maturation and intermediate progenitor maintenance during brain neurogenesis. PLoS Biol. 2010, 8, e1000460. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Sun, G.; Ye, P.; Murai, K.; Lang, M.F.; Li, S.; Zhang, H.; Li, W.; Fu, C.; Yin, J.; Wang, A.; et al. miR-137 forms a regulatory loop with nuclear receptor TLX and LSD1 in neural stem cells. Nat. Commun. 2011, 2, 529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Silber, J.; Lim, D.A.; Petritsch, C.; Persson, A.I.; Maunakea, A.K.; Yu, M.; Vandenberg, S.R.; Ginzinger, D.G.; James, C.D.; Costello, J.F.; et al. miR-124 and miR-137 inhibit proliferation of glioblastoma multiforme cells and induce differentiation of brain tumor stem cells. BMC Med. 2008, 6, 14. [Google Scholar] [CrossRef]
  44. Bemis, L.T.; Chen, R.; Amato, C.M.; Classen, E.H.; Robinson, S.E.; Coffey, D.G.; Erickson, P.F.; Shellman, Y.G.; Robinson, W.A. MicroRNA-137 targets microphthalmia-associated transcription factor in melanoma cell lines. Cancer Res. 2008, 68, 1362–1368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Wright, C.; Turner, J.A.; Calhoun, V.D.; Perrone-Bizzozero, N. Potential Impact of miR-137 and Its Targets in Schizophrenia. Front. Genet. 2013, 4, 58. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Szulwach, K.E.; Li, X.; Smrt, R.D.; Li, Y.; Luo, Y.; Lin, L.; Santistevan, N.J.; Li, W.; Zhao, X.; Jin, P. Cross talk between microRNA and epigenetic regulation in adult neurogenesis. J. Cell Biol. 2010, 189, 127–141. [Google Scholar] [CrossRef] [Green Version]
  47. Channakkar, A.S.; Singh, T.; Pattnaik, B.; Gupta, K.; Seth, P.; Adlakha, Y.K. MiRNA-137-mediated modulation of mitochondrial dynamics regulates human neural stem cell fate. Stem Cells 2020, 38, 683–697. [Google Scholar] [CrossRef] [Green Version]
  48. Smrt, R.D.; Szulwach, K.E.; Pfeiffer, R.L.; Li, X.; Guo, W.; Pathania, M.; Teng, Z.Q.; Luo, Y.; Peng, J.; Bordey, A.; et al. MicroRNA miR-137 regulates neuronal maturation by targeting ubiquitin ligase mind bomb-1. Stem Cells 2010, 28, 1060–1070. [Google Scholar] [CrossRef] [Green Version]
  49. Yamada, T.; Urano-Tashiro, Y.; Tanaka, S.; Akiyama, H.; Tashiro, F. Involvement of crosstalk between Oct4 and Meis1a in neural cell fate decision. PLoS ONE 2013, 8, e56997. [Google Scholar] [CrossRef] [Green Version]
  50. Mercurio, S.; Serra, L.; Nicolis, S.K. More than just Stem Cells: Functional Roles of the Transcription Factor Sox2 in Differentiated Glia and Neurons. Int. J. Mol. Sci. 2019, 20, 4540. [Google Scholar] [CrossRef]
  51. Boyer, L.A.; Lee, T.I.; Cole, M.F.; Johnstone, S.E.; Levine, S.S.; Zucker, J.P.; Guenther, M.G.; Kumar, R.M.; Murray, H.L.; Jenner, R.G.; et al. Core transcriptional regulatory circuitry in human embryonic stem cells. Cell 2005, 122, 947–956. [Google Scholar] [CrossRef] [Green Version]
  52. Mahmoudi, E.; Cairns, M.J. MiR-137: An important player in neural development and neoplastic transformation. Mol. Psychiatry 2017, 22, 44–55. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Tsuyama, J.; Bunt, J.; Richards, L.J.; Iwanari, H.; Mochizuki, Y.; Hamakubo, T.; Shimazaki, T.; Okano, H. MicroRNA-153 Regulates the Acquisition of Gliogenic Competence by Neural Stem Cells. Stem Cell Rep. 2015, 5, 365–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Wei, C.; Salichos, L.; Wittgrove, C.M.; Rokas, A.; Patton, J.G. Transcriptome-wide analysis of small RNA expression in early zebrafish development. RNA 2012, 18, 915–929. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Qiao, J.; Zhao, J.; Chang, S.; Sun, Q.; Liu, N.; Dong, J.; Chen, Y.; Yang, D.; Ye, D.; Liu, X.; et al. MicroRNA-153 improves the neurogenesis of neural stem cells and enhances the cognitive ability of aged mice through the notch signaling pathway. Cell Death Differ. 2020, 27, 808–825. [Google Scholar] [CrossRef] [PubMed]
  56. Bertoli, G.; Cava, C.; Castiglioni, I. MicroRNAs: New Biomarkers for Diagnosis, Prognosis, Therapy Prediction and Therapeutic Tools for Breast Cancer. Theranostics 2015, 5, 1122–1143. [Google Scholar] [CrossRef] [Green Version]
  57. Ishikawa, M.; Iwasaki, M.; Zhao, H.; Saito, J.; Hu, C.; Sun, Q.; Sakamoto, A.; Ma, D. Inhalational Anesthetics Inhibit Neuroglioma Cell Proliferation and Migration via miR-138, -210 and -335. Int. J. Mol. Sci. 2021, 22, 4355. [Google Scholar] [CrossRef]
  58. Gong, X.; Wang, Y.; Zeng, J.; Li, S.; Luo, Y. Computational identification and experimental validation of microRNAs binding to the fragile X syndrome gene Fmr1. Neurochem. Res. 2015, 40, 109–117. [Google Scholar] [CrossRef]
  59. Wu, C.; Zhang, X.; Chen, P.; Ruan, X.; Liu, W.; Li, Y.; Sun, C.; Hou, L.; Yin, B.; Qiang, B.; et al. MicroRNA-129 modulates neuronal migration by targeting Fmr1 in the developing mouse cortex. Cell Death Dis. 2019, 10, 287. [Google Scholar] [CrossRef] [Green Version]
  60. Liu, B.; Li, Y.; Stackpole, E.E.; Novak, A.; Gao, Y.; Zhao, Y.; Zhao, X.; Richter, J.D. Regulatory discrimination of mRNAs by FMRP controls mouse adult neural stem cell differentiation. Proc. Natl. Acad. Sci. USA 2018, 115, e11397–e11405. [Google Scholar] [CrossRef]
  61. Hou, L.; Liu, T.; Wang, J. Isoflurane suppresses the self-renewal of normal mouse neural stem cells in a p53-dependent manner by activating the Lkb1-p53-p21 signalling pathway. Mol. Med. Rep. 2015, 12, 7412–7418. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Wang, Z.; Park, H.J.; Carr, J.R.; Chen, Y.J.; Zheng, Y.; Li, J.; Tyner, A.L.; Costa, R.H.; Bagchi, S.; Raychaudhuri, P. FoxM1 in tumorigenicity of the neuroblastoma cells and renewal of the neural progenitors. Cancer Res. 2011, 71, 4292–4302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Jia, J.; Cui, Y.; Tan, Z.; Liu, M.; Jiang, Y. Transcriptional factor FoxM1-activated microRNA-335-3p maintains the self-renewal of neural stem cells by inhibiting p53 signaling pathway via Fmr1. Stem. Cell Res. Ther. 2021, 12, 169. [Google Scholar] [CrossRef] [PubMed]
  64. Xu, L.; Shen, J.; Yu, L.; Sun, J.; Yan, M. Autophagy is involved in sevoflurane-induced developmental neurotoxicity in the developing rat brain. Brain Res. Bull. 2018, 140, 226–232. [Google Scholar] [CrossRef]
  65. Shen, F.Y.; Song, Y.C.; Guo, F.; Xu, Z.D.; Li, Q.; Zhang, B.; Ma, Y.Q.; Zhang, Y.Q.; Lin, R.; Li, Y.; et al. Cognitive Impairment and Endoplasmic Reticulum Stress Induced by Repeated Short-Term Sevoflurane Exposure in Early Life of Rats. Front. Psychiatry 2018, 9, 332. [Google Scholar] [CrossRef] [Green Version]
  66. Pearn, M.L.; Schilling, J.M.; Jian, M.; Egawa, J.; Wu, C.; Mandyam, C.D.; Fannon-Pavlich, M.J.; Nguyen, U.; Bertoglio, J.; Kodama, M.; et al. Inhibition of RhoA reduces propofol-mediated growth cone collapse, axonal transport impairment, loss of synaptic connectivity, and behavioural deficits. Br. J. Anaesth. 2018, 120, 745–760. [Google Scholar] [CrossRef] [Green Version]
  67. Wang, Y.; Yin, S.; Xue, H.; Yang, Y.; Zhang, N.; Zhao, P. Mid-gestational sevoflurane exposure inhibits fetal neural stem cell proliferation and impairs postnatal learning and memory function in a dose-dependent manner. Dev. Biol. 2018, 435, 185–197. [Google Scholar] [CrossRef]
  68. Li, X.; Wu, Z.; Zhang, Y.; Xu, Y.; Han, G.; Zhao, P. Activation of Autophagy Contributes to Sevoflurane-Induced Neurotoxicity in Fetal Rats. Front. Mol. Neurosci. 2017, 10, 432. [Google Scholar] [CrossRef] [Green Version]
  69. Wang, Y.; Yin, S.W.; Zhang, N.; Zhao, P. High-concentration sevoflurane exposure in mid-gestation induces apoptosis of neural stem cells in rat offspring. Neural Regen. Res. 2018, 13, 1575–1584. [Google Scholar] [CrossRef]
  70. Schilling, G.; Wood, J.D.; Duan, K.; Slunt, H.H.; Gonzales, V.; Yamada, M.; Cooper, J.K.; Margolis, R.L.; Jenkins, N.A.; Copeland, N.G.; et al. Nuclear accumulation of truncated atrophin-1 fragments in a transgenic mouse model of DRPLA. Neuron 1999, 24, 275–286. [Google Scholar] [CrossRef]
  71. Zhang, F.; Xu, D.; Yuan, L.; Sun, Y.; Xu, Z. Epigenetic regulation of Atrophin1 by lysine-specific demethylase 1 is required for cortical progenitor maintenance. Nat. Commun. 2014, 5, 5815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Zhang, Y.; Wu, Z.; Li, X.; Wan, Y.; Zhang, Y.; Zhao, P. Maternal sevoflurane exposure affects differentiation of hippocampal neural stem cells by regulating miR-410-3p and ATN1. Stem. Cell Res. 2020, 11, 423. [Google Scholar] [CrossRef] [PubMed]
  73. Morgado, A.L.; Rodrigues, C.M.; Sola, S. MicroRNA-145 Regulates Neural Stem Cell Differentiation Through the Sox2-Lin28/let-7 Signaling Pathway. Stem. Cells 2016, 34, 1386–1395. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Cimadamore, F.; Amador-Arjona, A.; Chen, C.; Huang, C.T.; Terskikh, A.V. SOX2-LIN28/let-7 pathway regulates proliferation and neurogenesis in neural precursors. Proc. Natl. Acad. Sci. USA 2013, 110, E3017–E3026. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Takahashi, K.; Tanabe, K.; Ohnuki, M.; Narita, M.; Ichisaka, T.; Tomoda, K.; Yamanaka, S. Induction of pluripotent stem cells from adult human fibroblasts by defined factors. Cell 2007, 131, 861–872. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  76. Shi, J.; Li, H.; Jin, G.; Zhu, P.; Tian, M.; Qin, J.; Tan, X.; Zhao, S.; Wang, F.; Hua, Y.; et al. Lhx8 promote differentiation of hippocampal neural stem/progenitor cells into cholinergic neurons in vitro. Vitr. Cell Dev. Biol. Anim. 2012, 48, 603–609. [Google Scholar] [CrossRef]
  77. Cheng, X.; Li, W.; Zhao, R.; Li, H.; Qin, J.; Tian, M.; Zhang, X.; Jin, G. The role of hippocampal niche exosomes in rat hippocampal neurogenesis after fimbria-fornix transection. J. Biol. Chem. 2021, 296, 100188. [Google Scholar] [CrossRef]
  78. Li, W.; Wang, S.S.; Shan, B.Q.; Qin, J.B.; Zhao, H.Y.; Tian, M.L.; He, H.; Cheng, X.; Zhang, X.H.; Jin, G.H. miR-103-3p targets Ndel1 to regulate neural stem cell proliferation and differentiation. Neural Regen. Res. 2022, 17, 401–408. [Google Scholar] [CrossRef]
  79. Li, W.; Shan, B.Q.; Zhao, H.Y.; He, H.; Tian, M.L.; Cheng, X.; Qin, J.B.; Jin, G.H. MiR-130a-3p regulates neural stem cell differentiation in vitro by targeting Acsl4. J. Cell Mol. Med. 2022, 26, 2717–2727. [Google Scholar] [CrossRef]
  80. Li, W.; Wang, S.; He, H.; Qin, J.; Cheng, X.; Zhao, H.; Tian, M.; Zhang, X.; Jin, G. Expression and function of Ndel1 during the differentiation of neural stem cells induced by hippocampal exosomesticle. Stem. Cell Res. Ther. 2021, 12, 51. [Google Scholar] [CrossRef]
  81. Yang, Q.; Wan, L.; Xiao, C.; Hu, H.; Wang, L.; Zhao, J.; Lei, Z.; Zhang, H.T. Inhibition of LHX2 by miR-124 suppresses cellular migration and invasion in non-small cell lung cancer. Oncol Lett 2017, 14, 3429–3436. [Google Scholar] [CrossRef] [Green Version]
  82. Chen, D.; Liu, J.; Wu, Z.; Li, S.H. Role of miR-132/methyl-CpG-binding protein 2 in the regulation of neural stem cell differentiation. Neural Regen. Res. 2021, 16, 345–349. [Google Scholar] [CrossRef]
  83. Deng, Z.; Wei, Y.; Yao, Y.; Gao, S.; Wang, X. Let-7f promotes the differentiation of neural stem cells in rats. Am. J. Transl. Res. 2020, 12, 5752–5761. [Google Scholar] [PubMed]
  84. Denli, A.M.; Cao, X.; Gage, F.H. miR-9 and TLX: Chasing tails in neural stem cells. Nat. Struct. Mol. Biol. 2009, 16, 346–347. [Google Scholar] [CrossRef] [PubMed]
  85. Zhao, C.; Sun, G.; Li, S.; Lang, M.F.; Yang, S.; Li, W.; Shi, Y. MicroRNA let-7b regulates neural stem cell proliferation and differentiation by targeting nuclear receptor TLX signaling. Proc. Natl. Acad. Sci. USA 2010, 107, 1876–1881. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  86. Liu, C.; Teng, Z.Q.; Santistevan, N.J.; Szulwach, K.E.; Guo, W.; Jin, P.; Zhao, X. Epigenetic regulation of miR-184 by MBD1 governs neural stem cell proliferation and differentiation. Cell Stem. Cell 2010, 6, 433–444. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Guo, F.; Han, X.; Zhang, J.; Zhao, X.; Lou, J.; Chen, H.; Huang, X. Repetitive transcranial magnetic stimulation promotes neural stem cell proliferation via the regulation of MiR-25 in a rat model of focal cerebral ischemia. PLoS ONE 2014, 9, e109267. [Google Scholar] [CrossRef] [Green Version]
  88. Jin, J.; Kim, S.N.; Liu, X.; Zhang, H.; Zhang, C.; Seo, J.S.; Kim, Y.; Sun, T. miR-17-92 Cluster Regulates Adult Hippocampal Neurogenesis, Anxiety, and Depression. Cell Rep. 2016, 16, 1653–1663. [Google Scholar] [CrossRef] [Green Version]
  89. Liu, C.; Teng, Z.Q.; McQuate, A.L.; Jobe, E.M.; Christ, C.C.; von Hoyningen-Huene, S.J.; Reyes, M.D.; Polich, E.D.; Xing, Y.; Li, Y.; et al. An epigenetic feedback regulatory loop involving microRNA-195 and MBD1 governs neural stem cell differentiation. PLoS ONE 2013, 8, e51436. [Google Scholar] [CrossRef] [Green Version]
  90. Han, J.; Kim, H.J.; Schafer, S.T.; Paquola, A.; Clemenson, G.D.; Toda, T.; Oh, J.; Pankonin, A.R.; Lee, B.S.; Johnston, S.T.; et al. Functional Implications of miR-19 in the Migration of Newborn Neurons in the Adult Brain. Neuron 2016, 91, 79–89. [Google Scholar] [CrossRef]
  91. Hu, J.J.; Qin, L.J.; Liu, Z.Y.; Liu, P.; Wei, H.P.; Wang, H.Y.; Zhao, C.C.; Ge, Z.M. miR-15a regulates oxygen glucose deprivation/reperfusion (OGD/R)-induced neuronal injury by targeting BDNF. Kaohsiung J. Med. Sci. 2020, 36, 27–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. Dykes, I.M.; Emanueli, C. Transcriptional and Post-transcriptional Gene Regulation by Long Non-coding RNA. Genom. Proteom. Bioinform. 2017, 15, 177–186. [Google Scholar] [CrossRef] [PubMed]
  93. Ferrè, F.; Colantoni, A.; Helmer-Citterich, M. Revealing protein-lncRNA interaction. Brief. Bioinform. 2016, 17, 106–116. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Cesana, M.; Cacchiarelli, D.; Legnini, I.; Santini, T.; Sthandier, O.; Chinappi, M.; Tramontano, A.; Bozzoni, I. A long noncoding RNA controls muscle differentiation by functioning as a competing endogenous RNA. Cell 2011, 147, 358–369. [Google Scholar] [CrossRef] [Green Version]
  95. Tay, Y.; Rinn, J.; Pandolfi, P.P. The multilayered complexity of ceRNA crosstalk and competition. Nature 2014, 505, 344–352. [Google Scholar] [CrossRef] [Green Version]
  96. Sun, Q.; Hao, Q.; Prasanth, K.V. Nuclear Long Noncoding RNAs: Key Regulators of Gene Expression. Trends Genet. TIG 2018, 34, 142–157. [Google Scholar] [CrossRef]
  97. Yang, Y.; Ren, M.; Song, C.; Li, D.; Soomro, S.H.; Xiong, Y.; Zhang, H.; Fu, H. LINC00461, a long non-coding RNA, is important for the proliferation and migration of glioma cells. Oncotarget 2017, 8, 84123–84139. [Google Scholar] [CrossRef] [Green Version]
  98. Zhang, L.; Yan, J.; Liu, Q.; Xie, Z.; Jiang, H. LncRNA Rik-203 contributes to anesthesia neurotoxicity via microRNA-101a-3p and GSK-3β-mediated neural differentiation. Sci. Rep. 2019, 9, 6822. [Google Scholar] [CrossRef] [Green Version]
  99. Zhang, L.; Xue, Z.; Yan, J.; Wang, J.; Liu, Q.; Jiang, H. LncRNA Riken-201 and Riken-203 modulates neural development by regulating the Sox6 through sequestering miRNAs. Cell Prolif. 2019, 52, e12573. [Google Scholar] [CrossRef]
  100. Panman, L.; Papathanou, M.; Laguna, A.; Oosterveen, T.; Volakakis, N.; Acampora, D.; Kurtsdotter, I.; Yoshitake, T.; Kehr, J.; Joodmardi, E.; et al. Sox6 and Otx2 control the specification of substantia nigra and ventral tegmental area dopamine neurons. Cell Rep. 2014, 8, 1018–1025. [Google Scholar] [CrossRef]
  101. Azim, E.; Jabaudon, D.; Fame, R.M.; Macklis, J.D. SOX6 controls dorsal progenitor identity and interneuron diversity during neocortical development. Nat. Neurosci. 2009, 12, 1238–1247. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  102. Zhang, L.; Xue, Z.; Liu, Q.; Liu, Y.; Xi, S.; Cheng, Y.; Li, J.; Yan, J.; Shen, Y.; Xiao, C.; et al. Disrupted folate metabolism with anesthesia leads to myelination deficits mediated by epigenetic regulation of ERMN. EBioMedicine 2019, 43, 473–486. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  103. Jiang, Y.; Wang, Y.; Sun, Y.; Jiang, H. Long non-coding RNA Peg13 attenuates the sevoflurane toxicity against neural stem cells by sponging microRNA-128-3p to preserve Sox13 expression. PLoS ONE 2020, 15, e0243644. [Google Scholar] [CrossRef] [PubMed]
  104. Zhang, L.; Xue, Z.; Yan, J.; Jiang, H. LncRNA Rik-203 Contributes to Sevoflurane Induced Neurotoxicity? Front. Med. 2020, 7, 353. [Google Scholar] [CrossRef] [PubMed]
  105. Zhang, J.; Yuan, L.; Zhang, X.; Hamblin, M.H.; Zhu, T.; Meng, F.; Li, Y.; Chen, Y.E.; Yin, K.J. Altered long non-coding RNA transcriptomic profiles in brain microvascular endothelium after cerebral ischemia. Exp. Neurol. 2016, 277, 162–170. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Keshavarz, M.; Tautz, D. The imprinted lncRNA Peg13 regulates sexual preference and the sex-specific brain transcriptome in mice. Proc. Natl. Acad. Sci. USA 2021, 118, e2022172118. [Google Scholar] [CrossRef]
  107. Feng, H.; Gui, Q.; Zhu, W.; Wu, G.; Dong, X.; Shen, M.; Luo, H.; Xue, S.; Cheng, Q. Long-noncoding RNA Peg13 alleviates epilepsy progression in mice via the miR-490-3p/Psmd11 axis to inactivate the Wnt/β-catenin pathway. Am. J. Transl. Res. 2020, 12, 7968–7981. [Google Scholar]
  108. Zhao, Y.; Liu, H.; Zhang, Q.; Zhang, Y. The functions of long non-coding RNAs in neural stem cell proliferation and differentiation. Cell Biosci. 2020, 10, 74. [Google Scholar] [CrossRef]
  109. Gao, H.; Zhang, Y.; Xue, H.; Zhang, Q.; Zhang, Y.; Shen, Y.; Bing, X. Long Non-coding RNA Peg13 Alleviates Hypoxic-Ischemic Brain Damage in Neonatal Mice via miR-20a-5p/XIAP Axis. Neurochem. Res. 2022, 47, 656–666. [Google Scholar] [CrossRef]
  110. Ramos, A.D.; Andersen, R.E.; Liu, S.J.; Nowakowski, T.J.; Hong, S.J.; Gertz, C.; Salinas, R.D.; Zarabi, H.; Kriegstein, A.R.; Lim, D.A. The long noncoding RNA Pnky regulates neuronal differentiation of embryonic and postnatal neural stem cells. Cell Stem. Cell 2015, 16, 439–447. [Google Scholar] [CrossRef] [Green Version]
  111. Grammatikakis, I.; Gorospe, M. Identification of neural stem cell differentiation repressor complex Pnky-PTBP1. Stem. Cell Investig. 2016, 3, 10. [Google Scholar] [CrossRef] [PubMed]
  112. Du, J.; Li, Y.; Su, Y.; Zhi, W.; Zhang, J.; Zhang, C.; Wang, J.; Deng, W.; Zhao, S. LncRNA Pnky Positively Regulates Neural Stem Cell Migration by Modulating mRNA Splicing and Export of Target Genes. Cell. Mol. Neurobiol. 2022, 48, 537–545. [Google Scholar] [CrossRef] [PubMed]
  113. Andersen, R.E.; Hong, S.J.; Lim, J.J.; Cui, M.; Harpur, B.A.; Hwang, E.; Delgado, R.N.; Ramos, A.D.; Liu, S.J.; Blencowe, B.J.; et al. The Long Noncoding RNA Pnky Is a Trans-acting Regulator of Cortical Development In Vivo. Dev. Cell 2019, 49, 632–642.e637. [Google Scholar] [CrossRef]
  114. Lin, B.; Lu, L.; Wang, Y.; Zhang, Q.; Wang, Z.; Cheng, G.; Duan, X.; Zhang, F.; Xie, M.; Le, H.; et al. Nanomedicine Directs Neuronal Differentiation of Neural Stem Cells via Silencing Long Noncoding RNA for Stroke Therapy. Nano Lett. 2021, 21, 806–815. [Google Scholar] [CrossRef] [PubMed]
  115. Cui, Y.; Yin, Y.; Xiao, Z.; Zhao, Y.; Chen, B.; Yang, B.; Xu, B.; Song, H.; Zou, Y.; Ma, X.; et al. LncRNA Neat1 mediates miR-124-induced activation of Wnt/β-catenin signaling in spinal cord neural progenitor cells. Stem Cell Res. Ther. 2019, 10, 400. [Google Scholar] [CrossRef]
  116. Geng, W.; Guo, X.; Zhang, L.; Ma, Y.; Wang, L.; Liu, Z.; Ji, H.; Xiong, Y. Resveratrol inhibits proliferation, migration and invasion of multiple myeloma cells via NEAT1-mediated Wnt/β-catenin signaling pathway. Biomed. Pharmacother. 2018, 107, 484–494. [Google Scholar] [CrossRef]
  117. Marchetti, B. Wnt/β-Catenin Signaling Pathway Governs a Full Program for Dopaminergic Neuron Survival, Neurorescue and Regeneration in the MPTP Mouse Model of Parkinson’s Disease. Int. J. Mol. Sci. 2018, 19, 3743. [Google Scholar] [CrossRef] [Green Version]
  118. Han, K.; Kang, N.; Yu, X.; Lu, J.; Ma, Y. lncRNA NEAT1-let 7b-P21 axis mediates the proliferation of neural stem cells cultured in vitro promoted by radial extracorporeal shock wave. Regen. Ther. 2022, 21, 139–147. [Google Scholar] [CrossRef]
  119. Katsel, P.; Roussos, P.; Fam, P.; Khan, S.; Tan, W.; Hirose, T.; Nakagawa, S.; Pletnikov, M.V.; Haroutunian, V. The expression of long noncoding RNA NEAT1 is reduced in schizophrenia and modulates oligodendrocytes transcription. NPJ Schizophr. 2019, 5, 3. [Google Scholar] [CrossRef] [Green Version]
  120. Ni, W.; Yao, S.; Zhou, Y.; Liu, Y.; Huang, P.; Zhou, A.; Liu, J.; Che, L.; Li, J. Long noncoding RNA GAS5 inhibits progression of colorectal cancer by interacting with and triggering YAP phosphorylation and degradation and is negatively regulated by the m(6)A reader YTHDF3. Mol. Cancer 2019, 18, 143. [Google Scholar] [CrossRef]
  121. Chen, X.; Ren, G.; Li, Y.; Chao, W.; Chen, S.; Li, X.; Xue, S. Level of LncRNA GAS5 and Hippocampal Volume are Associated with the Progression of Alzheimer’s Disease. Clin. Interv. Aging 2022, 17, 745–753. [Google Scholar] [CrossRef] [PubMed]
  122. Zhao, H.; Jin, T.; Cheng, X.; Qin, J.; Zhang, L.; He, H.; Xue, J.; Jin, G. GAS5 which is regulated by Lhx8 promotes the recovery of learning and memory in rats with cholinergic nerve injury. Life Sci. 2020, 260, 118388. [Google Scholar] [CrossRef] [PubMed]
  123. Knauss, J.L.; Miao, N.; Kim, S.N.; Nie, Y.; Shi, Y.; Wu, T.; Pinto, H.B.; Donohoe, M.E.; Sun, T. Long noncoding RNA Sox2ot and transcription factor YY1 co-regulate the differentiation of cortical neural progenitors by repressing Sox2. Cell Death Dis. 2018, 9, 799. [Google Scholar] [CrossRef] [Green Version]
  124. Ng, S.Y.; Bogu, G.K.; Soh, B.S.; Stanton, L.W. The long noncoding RNA RMST interacts with SOX2 to regulate neurogenesis. Mol. Cell 2013, 51, 349–359. [Google Scholar] [CrossRef] [Green Version]
  125. Ma, X.; Wang, Y.; Yin, H.; Hua, L.; Zhang, X.; Xiao, J.; Yuan, Q.; Wang, S.; Liu, Y.; Zhang, S.; et al. Down-regulated long non-coding RNA RMST ameliorates dopaminergic neuron damage in Parkinson’s disease rats via regulation of TLR/NF-kappaB signaling pathway. Brain Res. Bull 2021, 174, 22–30. [Google Scholar] [CrossRef]
  126. Li, W.; Shen, W.; Zhang, B.; Tian, K.; Li, Y.; Mu, L.; Luo, Z.; Zhong, X.; Wu, X.; Liu, Y.; et al. Long non-coding RNA LncKdm2b regulates cortical neuronal differentiation by cis-activating Kdm2b. Protein Cell 2020, 11, 161–186. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  127. Singer, R.A.; Arnes, L.; Cui, Y.; Wang, J.; Gao, Y.; Guney, M.A.; Burnum-Johnson, K.E.; Rabadan, R.; Ansong, C.; Orr, G.; et al. The Long Noncoding RNA Paupar Modulates PAX6 Regulatory Activities to Promote Alpha Cell Development and Function. Cell Metab. 2019, 30, 1091–1106.e8. [Google Scholar] [CrossRef]
  128. Cheng, X.; Li, H.; Zhao, H.; Li, W.; Qin, J.; Jin, G. Function and mechanism of long non-coding RNA Gm21284 in the development of hippocampal cholinergic neurons. Cell Biosci. 2019, 9, 72. [Google Scholar] [CrossRef] [Green Version]
  129. Weng, R.; Lu, C.; Liu, X.; Li, G.; Lan, Y.; Qiao, J.; Bai, M.; Wang, Z.; Guo, X.; Ye, D.; et al. Long Noncoding RNA-1604 Orchestrates Neural Differentiation through the miR-200c/ZEB Axis. Stem. Cells 2018, 36, 325–336. [Google Scholar] [CrossRef] [Green Version]
  130. Chen, L.; Feng, P.; Zhu, X.; He, S.; Duan, J.; Zhou, D. Long non-coding RNA Malat1 promotes neurite outgrowth through activation of ERK/MAPK signalling pathway in N2a cells. J. Cell Mol. Med. 2016, 20, 2102–2110. [Google Scholar] [CrossRef] [Green Version]
  131. Winzi, M.; Casas Vila, N.; Paszkowski-Rogacz, M.; Ding, L.; Noack, S.; Theis, M.; Butter, F.; Buchholz, F. The long noncoding RNA lncR492 inhibits neural differentiation of murine embryonic stem cells. PLoS ONE 2018, 13, e0191682. [Google Scholar] [CrossRef] [PubMed]
  132. Zheng, X.; Lin, C.; Li, Y.; Ye, J.; Zhou, J.; Guo, P. Long noncoding RNA BDNF-AS regulates ketamine-induced neurotoxicity in neural stem cell derived neurons. Biomed. Pharmacother. 2016, 82, 722–728. [Google Scholar] [CrossRef]
  133. Prajapati, B.; Fatma, M.; Maddhesiya, P.; Sodhi, M.K.; Fatima, M.; Dargar, T.; Bhagat, R.; Seth, P.; Sinha, S. Identification and epigenetic analysis of divergent long non-coding RNAs in multilineage differentiation of human Neural Progenitor Cells. RNA Biol. 2019, 16, 13–24. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  134. Dong, X.; Chen, K.; Cuevas-Diaz Duran, R.; You, Y.; Sloan, S.A.; Zhang, Y.; Zong, S.; Cao, Q.; Barres, B.A.; Wu, J.Q. Comprehensive Identification of Long Non-coding RNAs in Purified Cell Types from the Brain Reveals Functional LncRNA in OPC Fate Determination. PLoS Genet. 2015, 11, e1005669. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  135. He, D.; Wang, J.; Lu, Y.; Deng, Y.; Zhao, C.; Xu, L.; Chen, Y.; Hu, Y.C.; Zhou, W.; Lu, Q.R. lncRNA Functional Networks in Oligodendrocytes Reveal Stage-Specific Myelination Control by an lncOL1/Suz12 Complex in the CNS. Neuron 2017, 93, 362–378. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  136. Li, Y.; Guo, B.; Yang, R.; Xiao, Z.; Gao, X.; Yu, J.; Li, S.; Luo, Y. A novel long noncoding RNA lnc158 promotes the differentiation of mouse neural precursor cells into oligodendrocytes by targeting nuclear factor-IB. Neuroreport 2018, 29, 1121–1128. [Google Scholar] [CrossRef]
  137. Kasuga, Y.; Fudge, A.D.; Zhang, Y.; Li, H. Characterization of a long noncoding RNA Pcdh17it as a novel marker for immature premyelinating oligodendrocytes. Glia 2019, 67, 2166–2177. [Google Scholar] [CrossRef]
  138. Mills, J.D.; Kavanagh, T.; Kim, W.S.; Chen, B.J.; Waters, P.D.; Halliday, G.M.; Janitz, M. High expression of long intervening non-coding RNA OLMALINC in the human cortical white matter is associated with regulation of oligodendrocyte maturation. Mol. Brain 2015, 8, 2. [Google Scholar] [CrossRef] [Green Version]
  139. You, D.; You, H. Repression of long non-coding RNA MEG3 restores nerve growth and alleviates neurological impairment after cerebral ischemia-reperfusion injury in a rat model. Biomed. Pharmacother. 2019, 111, 1447–1457. [Google Scholar] [CrossRef]
  140. Liu, Y.; Liu, C.; Zhang, A.; Yin, S.; Wang, T.; Wang, Y.; Wang, M.; Liu, Y.; Ying, Q.; Sun, J.; et al. Down-regulation of long non-coding RNA MEG3 suppresses osteogenic differentiation of periodontal ligament stem cells (PDLSCs) through miR-27a-3p/IGF1 axis in periodontitis. Aging 2019, 11, 5334–5350. [Google Scholar] [CrossRef]
  141. Wang, C.; Li, G.; Wu, Y.; Xi, J.; Kang, J. LincRNA1230 inhibits the differentiation of mouse ES cells towards neural progenitors. Sci. China. Life Sci. 2016, 59, 443–454. [Google Scholar] [CrossRef] [PubMed]
  142. Zhao, C.G.; Qin, J.; Li, J.; Jiang, S.; Ju, F.; Sun, W.; Ren, Z.; Ji, Y.Q.; Wang, R.; Sun, X.L.; et al. LINGO-1 regulates Wnt5a signaling during neural stem and progenitor cell differentiation by modulating miR-15b-3p levels. Stem Cell Res. Ther. 2021, 12, 372. [Google Scholar] [CrossRef] [PubMed]
  143. Pamudurti, N.R.; Bartok, O.; Jens, M.; Ashwal-Fluss, R.; Stottmeister, C.; Ruhe, L.; Hanan, M.; Wyler, E.; Perez-Hernandez, D.; Ramberger, E.; et al. Translation of CircRNAs. Mol. Cell 2017, 66, 9–21.e27. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  144. Salzman, J. Circular RNA Expression: Its Potential Regulation and Function. Trends Genet. TIG 2016, 32, 309–316. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Rybak-Wolf, A.; Stottmeister, C.; Glažar, P.; Jens, M.; Pino, N.; Giusti, S.; Hanan, M.; Behm, M.; Bartok, O.; Ashwal-Fluss, R.; et al. Circular RNAs in the Mammalian Brain Are Highly Abundant, Conserved, and Dynamically Expressed. Mol. Cell 2015, 58, 870–885. [Google Scholar] [CrossRef] [Green Version]
  146. Huang, R.; Zhang, Y.; Han, B.; Bai, Y.; Zhou, R.; Gan, G.; Chao, J.; Hu, G.; Yao, H. Circular RNA HIPK2 regulates astrocyte activation via cooperation of autophagy and ER stress by targeting MIR124-2HG. Autophagy 2017, 13, 1722–1741. [Google Scholar] [CrossRef] [Green Version]
  147. Han, B.; Zhang, Y.; Zhang, Y.; Bai, Y.; Chen, X.; Huang, R.; Wu, F.; Leng, S.; Chao, J.; Zhang, J.H.; et al. Novel insight into circular RNA HECTD1 in astrocyte activation via autophagy by targeting MIR142-TIPARP: Implications for cerebral ischemic stroke. Autophagy 2018, 14, 1164–1184. [Google Scholar] [CrossRef] [Green Version]
  148. Lasda, E.; Parker, R. Circular RNAs: Diversity of form and function. RNA 2014, 20, 1829–1842. [Google Scholar] [CrossRef] [Green Version]
  149. Yang, G.; Zhang, Y.; Yang, J. Identification of Potentially Functional CircRNA-miRNA-mRNA Regulatory Network in Gastric Carcinoma using Bioinformatics Analysis. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2019, 25, 8777–8796. [Google Scholar] [CrossRef]
  150. Wang, G.; Han, B.; Shen, L.; Wu, S.; Yang, L.; Liao, J.; Wu, F.; Li, M.; Leng, S.; Zang, F.; et al. Silencing of circular RNA HIPK2 in neural stem cells enhances functional recovery following ischaemic stroke. EBioMedicine 2020, 52, 102660. [Google Scholar] [CrossRef] [Green Version]
  151. Fang, J.; Zhang, T.; Liu, Y.; Li, Y.; Zhou, S.; Song, D.; Zhao, Y.; Feng, R.; Zhang, X.; Li, L.; et al. PAX6 downregulates miR-124 expression to promote cell migration during embryonic stem cell differentiation. Stem Cells Dev. 2014, 23, 2297–2310. [Google Scholar] [CrossRef] [PubMed]
  152. Cheng, L.C.; Pastrana, E.; Tavazoie, M.; Doetsch, F. miR-124 regulates adult neurogenesis in the subventricular zone stem cell niche. Nat. Neurosci. 2009, 12, 399–408. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Murray-Stewart, T.; Sierra, J.C.; Piazuelo, M.B.; Mera, R.M.; Chaturvedi, R.; Bravo, L.E.; Correa, P.; Schneider, B.G.; Wilson, K.T.; Casero, R.A. Epigenetic silencing of miR-124 prevents spermine oxidase regulation: Implications for Helicobacter pylori-induced gastric cancer. Oncogene 2016, 35, 5480–5488. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  154. Cai, L.; Qi, B.; Wu, X.; Peng, S.; Zhou, G.; Wei, Y.; Xu, J.; Chen, S.; Liu, S. Circular RNA Ttc3 regulates cardiac function after myocardial infarction by sponging miR-15b. J. Mol. Cell. Cardiol. 2019, 130, 10–22. [Google Scholar] [CrossRef] [PubMed]
  155. Yu, L.; Wang, Q.; Liu, N.; Zhao, J.; Yu, J.; Tao, S. Circular RNA circ-Ttc3 protects HaCaT cells from hypoxic injury by downregulation of miR-449a. IUBMB Life 2020, 72, 505–514. [Google Scholar] [CrossRef] [PubMed]
  156. Ma, X.; Zhu, G.; Jiao, T.; Shao, F. Effects of circular RNA Ttc3/miR-148a/Rcan2 axis on inflammation and oxidative stress in rats with acute kidney injury induced by sepsis. Life Sci. 2021, 272, 119233. [Google Scholar] [CrossRef] [PubMed]
  157. 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]
  158. Ye, Y.; Xu, H.; Zhang, X.; Li, Z.; Jia, Y.; He, X.; Huang, J.H. Association between toll-like receptor 4 expression and neural stem cell proliferation in the hippocampus following traumatic brain injury in mice. Int. J. Mol. Sci. 2014, 15, 12651–12664. [Google Scholar] [CrossRef] [Green Version]
  159. Soupene, E.; Kuypers, F.A. ACBD6 protein controls acyl chain availability and specificity of the N-myristoylation modification of proteins. J. Lipid Res. 2019, 60, 624–635. [Google Scholar] [CrossRef] [Green Version]
  160. Soupene, E.; Serikov, V.; Kuypers, F.A. Characterization of an acyl-coenzyme A binding protein predominantly expressed in human primitive progenitor cells. J Lipid Res 2008, 49, 1103–1112. [Google Scholar] [CrossRef] [Green Version]
  161. Li, W.; Shan, B.; Cheng, X.; He, H.; Qin, J.; Zhao, H.; Tian, M.; Zhang, X.; Jin, G. circRNA Acbd6 promotes neural stem cell differentiation into cholinergic neurons via the miR-320-5p-Osbpl2 axis. J. Biol. Chem. 2022, 298, 101828. [Google Scholar] [CrossRef] [PubMed]
  162. Yang, M.; Xiang, G.; Yu, D.; Yang, G.; He, W.; Yang, S.; Zhou, G.; Liu, A. Hsa_circ_0002468 Regulates the Neuronal Differentiation of SH-SY5Y Cells by Modulating the MiR-561/E2F8 Axis. Med. Sci. Monit. Int. Med. J. Exp. Clin. Res. 2019, 25, 2511–2519. [Google Scholar] [CrossRef] [PubMed]
  163. Xu, X.; Gu, D.; Xu, B.; Yang, C.; Wang, L. Circular RNA circ_0005835 promotes promoted neural stem cells proliferation and differentiate to neuron and inhibits inflammatory cytokines levels through miR-576-3p in Alzheimer’s disease. Environ. Sci. Pollut. Res. Int. 2022, 29, 35934–35943. [Google Scholar] [CrossRef] [PubMed]
  164. Yilmaz, U. [Alzheimer’s disease]. Der Radiologe 2015, 55, 386–388. [Google Scholar] [CrossRef] [PubMed]
  165. Ballard, C.; Gauthier, S.; Corbett, A.; Brayne, C.; Aarsland, D.; Jones, E. Alzheimer’s disease. Lancet 2011, 377, 1019–1031. [Google Scholar] [CrossRef]
  166. Turk, A.; Kunej, T.; Peterlin, B. MicroRNA-Target Interaction Regulatory Network in Alzheimer’s Disease. J. Pers. Med. 2021, 11, 1275. [Google Scholar] [CrossRef]
  167. Pogue, A.I.; Lukiw, W.J. Up-regulated Pro-inflammatory MicroRNAs (miRNAs) in Alzheimer’s disease (AD) and Age-Related Macular Degeneration (AMD). Cell. Mol. Neurobiol. 2018, 38, 1021–1031. [Google Scholar] [CrossRef]
  168. Zhao, Y.; Pogue, A.I.; Lukiw, W.J. MicroRNA (miRNA) Signaling in the Human CNS in Sporadic Alzheimer’s Disease (AD)-Novel and Unique Pathological Features. Int. J. Mol. Sci. 2015, 16, 30105–30116. [Google Scholar] [CrossRef] [Green Version]
  169. Cervera-Carles, L.; Dols-Icardo, O.; Molina-Porcel, L.; Alcolea, D.; Cervantes-Gonzalez, A.; Muñoz-Llahuna, L.; Clarimon, J. Assessing circular RNAs in Alzheimer’s disease and frontotemporal lobar degeneration. Neurobiol. Aging 2020, 92, 7–11. [Google Scholar] [CrossRef]
  170. Dube, U.; Del-Aguila, J.L.; Li, Z.; Budde, J.P.; Jiang, S.; Hsu, S.; Ibanez, L.; Fernandez, M.V.; Farias, F.; Norton, J.; et al. An atlas of cortical circular RNA expression in Alzheimer disease brains demonstrates clinical and pathological associations. Nat. Neurosci. 2019, 22, 1903–1912. [Google Scholar] [CrossRef]
  171. Idda, M.L.; Munk, R.; Abdelmohsen, K.; Gorospe, M. Noncoding RNAs in Alzheimer’s disease. Wiley Interdiscip. Rev. RNA 2018, 9, e1463. [Google Scholar] [CrossRef]
  172. Jaber, V.R.; Zhao, Y.; Sharfman, N.M.; Li, W.; Lukiw, W.J. Addressing Alzheimer’s Disease (AD) Neuropathology Using Anti-microRNA (AM) Strategies. Mol. Neurobiol. 2019, 56, 8101–8108. [Google Scholar] [CrossRef] [PubMed]
  173. Wang, W.X.; Rajeev, B.W.; Stromberg, A.J.; Ren, N.; Tang, G.; Huang, Q.; Rigoutsos, I.; Nelson, P.T. The expression of microRNA miR-107 decreases early in Alzheimer’s disease and may accelerate disease progression through regulation of beta-site amyloid precursor protein-cleaving enzyme 1. J. Neurosci. Off. J. Soc. Neurosci. 2008, 28, 1213–1223. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  174. Nelson, P.T.; Wang, W.X. MiR-107 is reduced in Alzheimer’s disease brain neocortex: Validation study. J. Alzheimer’s Dis. JAD 2010, 21, 75–79. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Jahangard, Y.; Monfared, H.; Moradi, A.; Zare, M.; Mirnajafi-Zadeh, J.; Mowla, S.J. Therapeutic Effects of Transplanted Exosomes Containing miR-29b to a Rat Model of Alzheimer’s Disease. Front. Neurosci. 2020, 14, 564. [Google Scholar] [CrossRef] [PubMed]
  176. Hébert, S.S.; Horré, K.; Nicolaï, L.; Papadopoulou, A.S.; Mandemakers, W.; Silahtaroglu, A.N.; Kauppinen, S.; Delacourte, A.; De Strooper, B. Loss of microRNA cluster miR-29a/b-1 in sporadic Alzheimer’s disease correlates with increased BACE1/beta-secretase expression. Proc. Natl. Acad. Sci. USA 2008, 105, 6415–6420. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Souza, V.C.; Morais, G.S., Jr.; Henriques, A.D.; Machado-Silva, W.; Perez, D.I.V.; Brito, C.J.; Camargos, E.F.; Moraes, C.F.; Nóbrega, O.T. Whole-Blood Levels of MicroRNA-9 Are Decreased in Patients With Late-Onset Alzheimer Disease. Am. J. Alzheimer’s Dis. Other Dement. 2020, 35, 1533317520911573. [Google Scholar] [CrossRef]
  178. Zhang, J.; Hu, M.; Teng, Z.; Tang, Y.P.; Chen, C. Synaptic and cognitive improvements by inhibition of 2-AG metabolism are through upregulation of microRNA-188-3p in a mouse model of Alzheimer’s disease. J. Neurosci. Off. J. Soc. Neurosci. 2014, 34, 14919–14933. [Google Scholar] [CrossRef] [Green Version]
  179. Long, J.M.; Ray, B.; Lahiri, D.K. MicroRNA-339-5p down-regulates protein expression of β-site amyloid precursor protein-cleaving enzyme 1 (BACE1) in human primary brain cultures and is reduced in brain tissue specimens of Alzheimer disease subjects. J. Biol. Chem. 2014, 289, 5184–5198. [Google Scholar] [CrossRef] [Green Version]
  180. Su, D.; Chai, Y.; Yang, J.; Wang, X.; Liu, Y.; Ma, J.; Tang, X.; Mishra, C.; Chandra, S.R.; Yue, W.; et al. Lentivirus-Carried microRNA-195 Rescues Memory Deficits of Alzheimer’s Disease Transgenic Mouse by Attenuating the Generation of Amyloid Plaques. Front. Pharmacol. 2021, 12, 633805. [Google Scholar] [CrossRef]
  181. Zhu, H.C.; Wang, L.M.; Wang, M.; Song, B.; Tan, S.; Teng, J.F.; Duan, D.X. MicroRNA-195 downregulates Alzheimer’s disease amyloid-β production by targeting BACE1. Brain Res. Bull. 2012, 88, 596–601. [Google Scholar] [CrossRef] [PubMed]
  182. Kim, J.; Yoon, H.; Chung, D.E.; Brown, J.L.; Belmonte, K.C.; Kim, J. miR-186 is decreased in aged brain and suppresses BACE1 expression. J. Neurochem. 2016, 137, 436–445. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  183. Sala Frigerio, C.; Lau, P.; Salta, E.; Tournoy, J.; Bossers, K.; Vandenberghe, R.; Wallin, A.; Bjerke, M.; Zetterberg, H.; Blennow, K.; et al. Reduced expression of hsa-miR-27a-3p in CSF of patients with Alzheimer disease. Neurology 2013, 81, 2103–2106. [Google Scholar] [CrossRef]
  184. Dickson, J.R.; Kruse, C.; Montagna, D.R.; Finsen, B.; Wolfe, M.S. Alternative polyadenylation and miR-34 family members regulate tau expression. J. Neurochem. 2013, 127, 739–749. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  185. Jian, C.; Lu, M.; Zhang, Z.; Liu, L.; Li, X.; Huang, F.; Xu, N.; Qin, L.; Zhang, Q.; Zou, D. miR-34a knockout attenuates cognitive deficits in APP/PS1 mice through inhibition of the amyloidogenic processing of APP. Life Sci. 2017, 182, 104–111. [Google Scholar] [CrossRef]
  186. Delay, C.; Dorval, V.; Fok, A.; Grenier-Boley, B.; Lambert, J.C.; Hsiung, G.Y.; Hébert, S.S. MicroRNAs targeting Nicastrin regulate Aβ production and are affected by target site polymorphisms. Front. Mol. Neurosci. 2014, 7, 67. [Google Scholar] [CrossRef]
  187. Zhou, L.T.; Zhang, J.; Tan, L.; Huang, H.Z.; Zhou, Y.; Liu, Z.Q.; Lu, Y.; Zhu, L.Q.; Yao, C.; Liu, D. Elevated Levels of miR-144-3p Induce Cholinergic Degeneration by Impairing the Maturation of NGF in Alzheimer’s Disease. Front. Cell Dev. Biol. 2021, 9, 667412. [Google Scholar] [CrossRef]
  188. Cheng, C.; Li, W.; Zhang, Z.; Yoshimura, S.; Hao, Q.; Zhang, C.; Wang, Z. MicroRNA-144 is regulated by activator protein-1 (AP-1) and decreases expression of Alzheimer disease-related a disintegrin and metalloprotease 10 (ADAM10). J. Biol. Chem. 2013, 288, 13748–13761. [Google Scholar] [CrossRef] [Green Version]
  189. De, D.; Mukherjee, I.; Guha, S.; Paidi, R.K.; Chakrabarti, S.; Biswas, S.C.; Bhattacharyya, S.N. Rheb-mTOR activation rescues Aβ-induced cognitive impairment and memory function by restoring miR-146 activity in glial cells. Mol. Therapy. Nucleic Acids 2021, 24, 868–887. [Google Scholar] [CrossRef]
  190. Lukiw, W.J.; Andreeva, T.V.; Grigorenko, A.P.; Rogaev, E.I. Studying micro RNA Function and Dysfunction in Alzheimer’s Disease. Front. Genet. 2012, 3, 327. [Google Scholar] [CrossRef] [Green Version]
  191. Xu, D.; Sharma, C.; Hemler, M.E. Tetraspanin12 regulates ADAM10-dependent cleavage of amyloid precursor protein. FASEB J. 2009, 23, 3674–3681. [Google Scholar] [CrossRef] [PubMed]
  192. Liu, D.; Tang, H.; Li, X.Y.; Deng, M.F.; Wei, N.; Wang, X.; Zhou, Y.F.; Wang, D.Q.; Fu, P.; Wang, J.Z.; et al. Targeting the HDAC2/HNF-4A/miR-101b/AMPK Pathway Rescues Tauopathy and Dendritic Abnormalities in Alzheimer’s Disease. Mol. Ther. 2017, 25, 752–764. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Vilardo, E.; Barbato, C.; Ciotti, M.; Cogoni, C.; Ruberti, F. MicroRNA-101 regulates amyloid precursor protein expression in hippocampal neurons. J. Biol. Chem. 2010, 285, 18344–18351. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  194. Barbato, C.; Pezzola, S.; Caggiano, C.; Antonelli, M.; Frisone, P.; Ciotti, M.T.; Ruberti, F. A lentiviral sponge for miR-101 regulates RanBP9 expression and amyloid precursor protein metabolism in hippocampal neurons. Front. Cell. Neurosci. 2014, 8, 37. [Google Scholar] [CrossRef] [Green Version]
  195. Zhang, B.; Chen, C.F.; Wang, A.H.; Lin, Q.F. MiR-16 regulates cell death in Alzheimer’s disease by targeting amyloid precursor protein. Eur. Rev. Med. Pharmacol. Sci. 2015, 19, 4020–4027. [Google Scholar]
  196. Liu, W.; Liu, C.; Zhu, J.; Shu, P.; Yin, B.; Gong, Y.; Qiang, B.; Yuan, J.; Peng, X. MicroRNA-16 targets amyloid precursor protein to potentially modulate Alzheimer’s-associated pathogenesis in SAMP8 mice. Neurobiol. Aging 2012, 33, 522–534. [Google Scholar] [CrossRef]
  197. Delay, C.; Calon, F.; Mathews, P.; Hébert, S.S. Alzheimer-specific variants in the 3’UTR of Amyloid precursor protein affect microRNA function. Mol. Neurodegener. 2011, 6, 70. [Google Scholar] [CrossRef] [Green Version]
  198. An, F.; Gong, G.; Wang, Y.; Bian, M.; Yu, L.; Wei, C. MiR-124 acts as a target for Alzheimer’s disease by regulating BACE1. Oncotarget 2017, 8, 114065–114071. [Google Scholar] [CrossRef] [Green Version]
  199. Fang, M.; Wang, J.; Zhang, X.; Geng, Y.; Hu, Z.; Rudd, J.A.; Ling, S.; Chen, W.; Han, S. The miR-124 regulates the expression of BACE1/β-secretase correlated with cell death in Alzheimer’s disease. Toxicol. Lett. 2012, 209, 94–105. [Google Scholar] [CrossRef]
  200. Smith, P.; Al Hashimi, A.; Girard, J.; Delay, C.; Hébert, S.S. In vivo regulation of amyloid precursor protein neuronal splicing by microRNAs. J. Neurochem. 2011, 116, 240–247. [Google Scholar] [CrossRef]
  201. Deng, Y.; Zhang, J.; Sun, X.; Ma, G.; Luo, G.; Miao, Z.; Song, L. miR-132 improves the cognitive function of rats with Alzheimer’s disease by inhibiting the MAPK1 signal pathway. Exp. Ther. Med. 2020, 20, 159. [Google Scholar] [CrossRef] [PubMed]
  202. Smith, P.Y.; Hernandez-Rapp, J.; Jolivette, F.; Lecours, C.; Bisht, K.; Goupil, C.; Dorval, V.; Parsi, S.; Morin, F.; Planel, E.; et al. miR-132/212 deficiency impairs tau metabolism and promotes pathological aggregation in vivo. Hum. Mol. Genet. 2015, 24, 6721–6735. [Google Scholar] [CrossRef] [PubMed]
  203. Moncini, S.; Lunghi, M.; Valmadre, A.; Grasso, M.; Del Vescovo, V.; Riva, P.; Denti, M.A.; Venturin, M. The miR-15/107 Family of microRNA Genes Regulates CDK5R1/p35 with Implications for Alzheimer’s Disease Pathogenesis. Mol. Neurobiol. 2017, 54, 4329–4342. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  204. Hébert, S.S.; Papadopoulou, A.S.; Smith, P.; Galas, M.C.; Planel, E.; Silahtaroglu, A.N.; Sergeant, N.; Buée, L.; De Strooper, B. Genetic ablation of Dicer in adult forebrain neurons results in abnormal tau hyperphosphorylation and neurodegeneration. Hum. Mol. Genet. 2010, 19, 3959–3969. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  205. Jaouen, F.; Gascon, E. Understanding the Role of miR-33 in Brain Lipid Metabolism: Implications for Alzheimer’s Disease. J. Neurosci. Off. J. Soc. Neurosci. 2016, 36, 2558–2560. [Google Scholar] [CrossRef] [Green Version]
  206. Kim, J.; Yoon, H.; Horie, T.; Burchett, J.M.; Restivo, J.L.; Rotllan, N.; Ramírez, C.M.; Verghese, P.B.; Ihara, M.; Hoe, H.S.; et al. microRNA-33 Regulates ApoE Lipidation and Amyloid-β Metabolism in the Brain. J. Neurosci. Off. J. Soc. Neurosci. 2015, 35, 14717–14726. [Google Scholar] [CrossRef] [Green Version]
  207. Decourt, B.; Sabbagh, M.N. BACE1 as a potential biomarker for Alzheimer’s disease. J. Alzheimer’s Dis. JAD 2011, 24, 53–59. [Google Scholar] [CrossRef]
  208. Faghihi, M.A.; Modarresi, F.; Khalil, A.M.; Wood, D.E.; Sahagan, B.G.; Morgan, T.E.; Finch, C.E.; Laurent, G.S., 3rd; Kenny, P.J.; Wahlestedt, C. Expression of a noncoding RNA is elevated in Alzheimer’s disease and drives rapid feed-forward regulation of beta-secretase. Nat. Med. 2008, 14, 723–730. [Google Scholar] [CrossRef] [Green Version]
  209. Ciarlo, E.; Massone, S.; Penna, I.; Nizzari, M.; Gigoni, A.; Dieci, G.; Russo, C.; Florio, T.; Cancedda, R.; Pagano, A. An intronic ncRNA-dependent regulation of SORL1 expression affecting Aβ formation is upregulated in post-mortem Alzheimer’s disease brain samples. Dis. Model. Mech. 2013, 6, 424–433. [Google Scholar] [CrossRef] [Green Version]
  210. Feng, L.; Liao, Y.T.; He, J.C.; Xie, C.L.; Chen, S.Y.; Fan, H.H.; Su, Z.P.; Wang, Z. Plasma long non-coding RNA BACE1 as a novel biomarker for diagnosis of Alzheimer disease. BMC Neurol. 2018, 18, 4. [Google Scholar] [CrossRef] [Green Version]
  211. Modarresi, F.; Faghihi, M.A.; Lopez-Toledano, M.A.; Fatemi, R.P.; Magistri, M.; Brothers, S.P.; van der Brug, M.P.; Wahlestedt, C. Inhibition of natural antisense transcripts in vivo results in gene-specific transcriptional upregulation. Nature Biotechnol. 2012, 30, 453–459. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  212. Yamanaka, Y.; Faghihi, M.A.; Magistri, M.; Alvarez-Garcia, O.; Lotz, M.; Wahlestedt, C. Antisense RNA controls LRP1 Sense transcript expression through interaction with a chromatin-associated protein, HMGB2. Cell Rep. 2015, 11, 967–976. [Google Scholar] [CrossRef] [PubMed]
  213. Parenti, R.; Paratore, S.; Torrisi, A.; Cavallaro, S. A natural antisense transcript against Rad18, specifically expressed in neurons and upregulated during beta-amyloid-induced apoptosis. Eur. J. Neurosci. 2007, 26, 2444–2457. [Google Scholar] [CrossRef] [PubMed]
  214. Magistri, M.; Velmeshev, D.; Makhmutova, M.; Faghihi, M.A. Transcriptomics Profiling of Alzheimer’s Disease Reveal Neurovascular Defects, Altered Amyloid-β Homeostasis, and Deregulated Expression of Long Noncoding RNAs. J. Alzheimer’s Dis. JAD 2015, 48, 647–665. [Google Scholar] [CrossRef] [Green Version]
  215. Zhao, Y.; Alexandrov, P.N.; Jaber, V.; Lukiw, W.J. Deficiency in the Ubiquitin Conjugating Enzyme UBE2A in Alzheimer’s Disease (AD) is Linked to Deficits in a Natural Circular miRNA-7 Sponge (circRNA; ciRS-7). Genes 2016, 7, 116. [Google Scholar] [CrossRef] [Green Version]
  216. Shi, Z.; Chen, T.; Yao, Q.; Zheng, L.; Zhang, Z.; Wang, J.; Hu, Z.; Cui, H.; Han, Y.; Han, X.; et al. The circular RNA ciRS-7 promotes APP and BACE1 degradation in an NF-κB-dependent manner. FEBS J. 2017, 284, 1096–1109. [Google Scholar] [CrossRef] [Green Version]
  217. Peng, W.; Mao, L.; Yin, D.; Sun, W.; Wang, H.; Zhang, Q.; Wang, J.; Chen, C.; Zeng, M.; Ding, J.; et al. Functional network changes in the hippocampus contribute to depressive symptoms in epilepsy. Seizure 2018, 60, 16–22. [Google Scholar] [CrossRef] [Green Version]
  218. Szot, P. Common factors among Alzheimer’s disease, Parkinson’s disease, and epilepsy: Possible role of the noradrenergic nervous system. Epilepsia 2012, 53, 61–66. [Google Scholar] [CrossRef]
  219. Huberfeld, G.; Blauwblomme, T.; Miles, R. Hippocampus and epilepsy: Findings from human tissues. Rev. Neurol. 2015, 171, 236–251. [Google Scholar] [CrossRef] [Green Version]
  220. Doria, J.W.; Forgacs, P.B. Incidence, Implications, and Management of Seizures Following Ischemic and Hemorrhagic Stroke. Curr. Neurol. Neurosci. Rep. 2019, 19, 37. [Google Scholar] [CrossRef]
  221. Wang, J.Z.; Vyas, M.V.; Saposnik, G.; Burneo, J.G. Incidence and management of seizures after ischemic stroke: Systematic review and meta-analysis. Neurology 2017, 89, 1220–1228. [Google Scholar] [CrossRef] [PubMed]
  222. Zhang, C.; Wang, X.; Wang, Y.; Zhang, J.G.; Hu, W.; Ge, M.; Zhang, K.; Shao, X. Risk factors for post-stroke seizures: A systematic review and meta-analysis. Epilepsy Res. 2014, 108, 1806–1816. [Google Scholar] [CrossRef] [PubMed]
  223. Manna, I.; Fortunato, F.; De Benedittis, S.; Sammarra, I.; Bertoli, G.; Labate, A.; Gambardella, A. Non-Coding RNAs: New Biomarkers and Therapeutic Targets for Temporal Lobe Epilepsy. Int. J. Mol. Sci. 2022, 23, 3063. [Google Scholar] [CrossRef] [PubMed]
  224. Ghafouri-Fard, S.; Hussen, B.M.; Jamali, E.; Branicki, W.; Taheri, M.; Akbari Dilmaghani, N. Role of lncRNAs and circRNAs in epilepsy. Ageing Res. Rev. 2022, 82, 101749. [Google Scholar] [CrossRef] [PubMed]
  225. Gorter, J.A.; Iyer, A.; White, I.; Colzi, A.; van Vliet, E.A.; Sisodiya, S.; Aronica, E. Hippocampal subregion-specific microRNA expression during epileptogenesis in experimental temporal lobe epilepsy. Neurobiol. Dis. 2014, 62, 508–520. [Google Scholar] [CrossRef]
  226. Veno, M.T.; Reschke, C.R.; Morris, G.; Connolly, N.M.C.; Su, J.; Yan, Y.; Engel, T.; Jimenez-Mateos, E.M.; Harder, L.M.; Pultz, D.; et al. A systems approach delivers a functional microRNA catalog and expanded targets for seizure suppression in temporal lobe epilepsy. Proc. Natl. Acad. Sci. USA 2020, 117, 15977–15988. [Google Scholar] [CrossRef]
  227. Wang, W.; Wang, X.; Chen, L.; Zhang, Y.; Xu, Z.; Liu, J.; Jiang, G.; Li, J.; Zhang, X.; Wang, K.; et al. The microRNA miR-124 suppresses seizure activity and regulates CREB1 activity. Expert Rev. Mol. Med. 2016, 18, e4. [Google Scholar] [CrossRef]
  228. Wang, H.; Ye, Y.; Zhu, Z.; Mo, L.; Lin, C.; Wang, Q.; Wang, H.; Gong, X.; He, X.; Lu, G.; et al. MiR-124 Regulates Apoptosis and Autophagy Process in MPTP Model of Parkinson’s Disease by Targeting to Bim. Brain Pathol. 2016, 26, 167–176. [Google Scholar] [CrossRef]
  229. Jimenez-Mateos, E.M.; Engel, T.; Merino-Serrais, P.; Fernaud-Espinosa, I.; Rodriguez-Alvarez, N.; Reynolds, J.; Reschke, C.R.; Conroy, R.M.; McKiernan, R.C.; deFelipe, J.; et al. Antagomirs targeting microRNA-134 increase hippocampal pyramidal neuron spine volume in vivo and protect against pilocarpine-induced status epilepticus. Brain Struct. Funct. 2015, 220, 2387–2399. [Google Scholar] [CrossRef]
  230. Luo, H.C.; Yi, T.Z.; Huang, F.G.; Wei, Y.; Luo, X.P.; Luo, Q.S. Role of long noncoding RNA MEG3/miR-378/GRB2 axis in neuronal autophagy and neurological functional impairment in ischemic stroke. J. Biol. Chem. 2020, 295, 14125–14139. [Google Scholar] [CrossRef]
  231. Xin, H.; Katakowski, M.; Wang, F.; Qian, J.Y.; Liu, X.S.; Ali, M.M.; Buller, B.; Zhang, Z.G.; Chopp, M. MicroRNA cluster miR-17-92 Cluster in Exosomes Enhance Neuroplasticity and Functional Recovery After Stroke in Rats. Stroke 2017, 48, 747–753. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  232. Yang, X.; Tang, X.; Sun, P.; Shi, Y.; Liu, K.; Hassan, S.H.; Stetler, R.A.; Chen, J.; Yin, K.J. MicroRNA-15a/16-1 Antagomir Ameliorates Ischemic Brain Injury in Experimental Stroke. Stroke 2017, 48, 1941–1947. [Google Scholar] [CrossRef] [PubMed]
  233. Chu, B.; Zhou, Y.; Zhai, H.; Li, L.; Sun, L.; Li, Y. The role of microRNA-146a in regulating the expression of IRAK1 in cerebral ischemia-reperfusion injury. Can. J. Physiol. Pharmacol. 2018, 96, 611–617. [Google Scholar] [CrossRef] [PubMed]
  234. 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] [CrossRef]
  235. Xie, Y.L.; Zhang, B.; Jing, L. MiR-125b blocks Bax/Cytochrome C/Caspase-3 apoptotic signaling pathway in rat models of cerebral ischemia-reperfusion injury by targeting p53. Neurol. Res. 2018, 40, 828–837. [Google Scholar] [CrossRef]
  236. Zhang, J.F.; Shi, L.L.; Zhang, L.; Zhao, Z.H.; Liang, F.; Xu, X.; Zhao, L.Y.; Yang, P.B.; Zhang, J.S.; Tian, Y.F. MicroRNA-25 Negatively Regulates Cerebral Ischemia/Reperfusion Injury-Induced Cell Apoptosis Through Fas/FasL Pathway. J. Mol. Neurosci. 2016, 58, 507–516. [Google Scholar] [CrossRef]
  237. Wang, X.; Shi, C.; Pan, H.; Meng, X.; Ji, F. MicroRNA-22 exerts its neuroprotective and angiogenic functions via regulating PI3K/Akt signaling pathway in cerebral ischemia-reperfusion rats. J. Neural Transm. 2020, 127, 35–44. [Google Scholar] [CrossRef]
  238. Wan, P.; Su, W.; Zhang, Y.; Li, Z.; Deng, C.; Li, J.; Jiang, N.; Huang, S.; Long, E.; Zhuo, Y. LncRNA H19 initiates microglial pyroptosis and neuronal death in retinal ischemia/reperfusion injury. Cell Death Differ. 2020, 27, 176–191. [Google Scholar] [CrossRef] [Green Version]
  239. 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]
  240. Jing, H.; Liu, L.; Jia, Y.; Yao, H.; Ma, F. Overexpression of the long non-coding RNA Oprm1 alleviates apoptosis from cerebral ischemia-reperfusion injury through the Oprm1/miR-155/GATA3 axis. Artif. Cells Nanomed. Biotechnol. 2019, 47, 2431–2439. [Google Scholar] [CrossRef]
  241. Samidurai, A.; Ockaili, R.A.; Cain, C.; Roh, S.K.; Kraskauskas, D.; Salloum, F.N.; Kukreja, R.C.; Das, A. Abstract 16178: Role of LncRNA MALAT-1 and MicroRNA-199 in Rapamycin Induced Cardioprotection Against Reperfusion Injury in Diabetic Rabbits. Circulation 2019, 140, A16178. [Google Scholar]
Figure 1. The feedback loops of miRNAs regulated by MeCP2 promoting or repressing neuronal differentiation.
Figure 1. The feedback loops of miRNAs regulated by MeCP2 promoting or repressing neuronal differentiation.
Biomolecules 13 00018 g001
Table 4. Summary of ncRNAs involved in AD.
Table 4. Summary of ncRNAs involved in AD.
NameTarget Protein/Gene/ProcessReferences
miRNAs
miR-107BACE1[173,174]
miR-29a/b-1BACE1[175,176]
miR-9BACE1[176,177]
miR-188-3pBACE1[178]
miR-339-5pBACE1[179]
miR-195BACE1[180,181]
miR-186BACE1[182]
miR-27a-3pPresenilin-1[183]
miRNA-34aTau, γ-secretase[184,185]
miRs-24, miR-186, miR-455Nicastrin[186]
miR-144ADAM10[187,188]
miR-125b, miR-146aTetraspanin12[189,190,191]
miR-101APP, Tau[192,193,194]
miR-16APP[195,196]
miR-147, miR-20aAPP[197]
miR-124PTBP1, BACE1[198,199,200]
miR-132, miR-212GSK3, Tau, INOS[201,202]
miR-15, miR-16ERK1, CDK5R1/p35[203,204]
miR-33ABCA1 and APOE lipidation[205,206]
lncRNA
BACE1-ASBACE1[207,208]
lncRNA 51ASORL1[209,210]
BDNF-AS (NAT)BDNF[209,210,211]
LRP1-ASLRP1[210,212]
NAT-Rad18RAD18[210,213]
AD-linc1, AD-linc2Unknown[210,214]
EBF3-AS (NAT)Nuclear processes[210,214]
HAO2-AS (NAT)Nuclear processes[210,214]
circRNA
ciRS-7Sponge for miR-7 to regulate UBE2A[215,216]
Table 5. Summary of ncRNAs involved in stroke and I/R injury.
Table 5. Summary of ncRNAs involved in stroke and I/R injury.
NamePutative Regulators and TargetsSignaling PathwaysFunctionsRef.
miRNA
miR-146aIL-1β, IRAK1, TNF-α, IL-6NF-κBAggravate cerebral I/R injury[233]
miR-155-5pDUSP14NF-κB, MAPKAccelerate cerebral I/R injury[234]
miR-125bp53Bax/Cytochrome C/Caspase-3Protect cerebral I/R injury[235]
miR-25Bax, Caspase-3, Bcl-2Fas/FasLInhibit cerebral I/R injury-induced apoptosis[236]
miR-22Ang-1, VEGFPI3K/AKTProtect and promote angiogenesis in cerebral I/R[237]
lncRNA
H19Beclin1, DUSP5, LC3I, L3II, P62ERK1/2Onset of I/R-induced inflammation[238]
MEG3BDNF, bFGF, NGFWnt/β-cateninEnhance OGD/R-induced pyroptosis and inflammation during cerebral I/R injury[239]
Oprm1GATA3NF-κBAlleviate apoptosis in cerebral I/R injury[240]
MALAT1miR-145, AQP4 Accelerate cerebral I/R injury[241]
circRNA
circTTC3TLR4 Accelerate cerebral I/R injury[157]
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

Tan, Z.; Li, W.; Cheng, X.; Zhu, Q.; Zhang, X. Non-Coding RNAs in the Regulation of Hippocampal Neurogenesis and Potential Treatment Targets for Related Disorders. Biomolecules 2023, 13, 18. https://doi.org/10.3390/biom13010018

AMA Style

Tan Z, Li W, Cheng X, Zhu Q, Zhang X. Non-Coding RNAs in the Regulation of Hippocampal Neurogenesis and Potential Treatment Targets for Related Disorders. Biomolecules. 2023; 13(1):18. https://doi.org/10.3390/biom13010018

Chicago/Turabian Style

Tan, Zhengye, Wen Li, Xiang Cheng, Qing Zhu, and Xinhua Zhang. 2023. "Non-Coding RNAs in the Regulation of Hippocampal Neurogenesis and Potential Treatment Targets for Related Disorders" Biomolecules 13, no. 1: 18. https://doi.org/10.3390/biom13010018

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