Skip to main content

Emerging role of non-coding RNAs in neuroinflammation mediated by microglia and astrocytes

Abstract

Neuroinflammation has been implicated in the initiation and progression of several central nervous system (CNS) disorders, including Alzheimer’s disease, Parkinson’s disease, amyotrophic lateral sclerosis, multiple sclerosis, ischemic stroke, traumatic brain injury, spinal cord injury, viral encephalitis, and bacterial encephalitis. Microglia and astrocytes are essential in neural development, maintenance of synaptic connections, and homeostasis in a healthy brain. The activation of astrocytes and microglia is a defense mechanism of the brain against damaged tissues and harmful pathogens. However, their activation triggers neuroinflammation, which can exacerbate or induce CNS injury. Non-coding RNAs (ncRNAs) are functional RNA molecules that lack coding capabilities but can actively regulate mRNA expression and function through various mechanisms. ncRNAs are highly expressed in astrocytes and microglia and are potential mediators of neuroinflammation. We reviewed the recent research progress on the role of miRNAs, lncRNAs, and circRNAs in regulating neuroinflammation in various CNS diseases. Understanding how these ncRNAs affect neuroinflammation will provide important therapeutic insights for preventing and managing CNS dysfunction.

Introduction

The central nervous system (CNS) is considered an immune-privileged site compared to the peripheral tissues [1]. The lack of resident dendritic cells and the relatively anti-inflammatory environment of the neural tissue result in a muted innate immune response within the CNS parenchyma [2]. However, the immune response in the CNS is widespread, and CNS homeostasis is highly dependent on the balance of the innate immune response. Neuroinflammation is a complex innate immune response involving reactive CNS elements that alter homeostasis. Inflammation is mediated by proinflammatory cytokines, chemokines, reactive oxygen species, and secondary messengers. Initially, neuroinflammation plays a beneficial role by eliminating microbes or promoting tissue repair. Conversely, uncontrolled neuroinflammation can become detrimental and cause pathogenic tissue damage [3]. Therefore, understanding the cellular and molecular regulators of neuroinflammation may provide helpful clues for developing new therapeutic interventions for treating CNS diseases.

Glial cells are critical components of the CNS and include microglia, astrocytes, and oligodendrocytes [4]. Microglia are the resident phagocytes of the innate immune system and the most motile cells in the CNS that regulate brain development, maintain neuronal networks, and modulate CNS injury and infection [5]. Microglia are activated earlier than other glial cells and are first responders to various CNS insults. Microglial activation is accompanied by morphological changes and is categorized into two opposing types: M1 and M2 phenotypes [6]. M1 microglia exhibit proinflammatory and neurotoxic states and are involved in the acute defense against pathogenic organisms, whereas M2 microglia are involved in the resolution of inflammation and tissue repair. Astrocytes are the most abundant type of glial cells in the CNS and are essential for brain homeostasis. Astrocytes provide an energy substrate for neurons, maintain the extracellular balance of ions and fluid, and contribute to the formation and maintenance of the blood–brain barrier (BBB) [7]. In addition, astrocytes, which are immune-competent cells within the brain, are critical regulators of innate and adaptive immune responses in the injured CNS. Upon CNS insult, astrocytes undergo proliferation and morphological changes, termed astrogliosis [8]. Increased expression of the glial fibrillary acidic protein (GFAP) is a marker of astrogliosis. Astrocytes can exacerbate inflammatory responses and aggravate tissue damage, but they can also promote immunosuppression and tissue repair. Thus, the specific role of glial cells depends on their unique characteristics and the nature of stimuli present in the inflammatory environment [9].

High-throughput sequencing techniques have revealed that only 1–2% of the human genome encodes proteins, but up to 90% of genome-produced transcripts have no protein-coding capacity and are referred to as non-coding RNAs (ncRNAs) [10]. ncRNAs are divided into two subclasses based on their biological functions: housekeeping and regulatory ncRNAs. Housekeeping ncRNAs (tRNA, rRNA, etc.) are constitutively and ubiquitously expressed, and essential for cell maintenance [11]. The three distinct classes of regulatory ncRNAs are microRNAs (miRNAs), long non-coding RNAs (lncRNAs), and circular RNAs (circRNAs) [12]. miRNAs are a large family of short, single-stranded ncRNAs, approximately 22 nucleotides (nt) in length. They play an important role in post-transcriptional gene regulation by targeting 3ʹ-untranslated regions (UTRs), resulting in translational repression or degradation of their messenger RNA (mRNA) targets [13]. Studies indicate that miRNAs are also capable of translation promotion and DNA binding for repression or activity [14, 15]. Importantly, at least 60% of protein-coding genes are targeted by miRNAs, leading to their broad regulatory roles in diverse biological processes [16]. lncRNAs are defined as transcripts longer than 200 nt without coding capacity. Based on their genomic locations, lncRNAs are classified into four different types: long intergenicncRNAs, antisense, intronic, and bidirectional lncRNAs [17]. In contrast to miRNAs, lncRNAs have been proposed to have regulatory functions in gene expression at both the transcriptional and post-transcriptional levels in various cellular and biological processes [18]. They regulate the expression of genes located on the same chromosome (acting in cis) or genes from different chromosomes (acting in trans) by interacting with proteins, DNAs, or RNAs. circRNAs are covalently bound endogenous molecules generated by a non-canonical splicing event called back-splicing [19]. During back-splicing, a downstream splice donor site is linked to an upstream splice acceptor site, and a phosphodiester bond at the junction site ligates the RNA cycle [20]. circRNAs are molecules with high stability due to the absence of 5ʹ caps and 3ʹ poly-A tails, and the covalently closed ring structure protects against exonuclease-mediated degradation [21]. In addition, circRNAs exhibit cell- and tissue-specific patterns and are particularly abundant in the human brain [22]. Recent investigations have shown that circRNAs are involved in neuronal function, cell proliferation, and innate immunity. At the molecular level, circRNAs modulate gene expression by sponging miRNAs, interacting with proteins, and regulating transcription and splicing [23].ncRNAs are differentially expressed in various neurological disorders associated with neuroinflammation, including chronic neurodegenerative diseases, acute neurodegenerative diseases, and CNS infectious diseases. In addition, microglia and astrocytes are closely associated with the outcome and progression of CNS disorders. The emerging links between ncRNAs, glial cells, and CNS disorders have opened a new field of diagnostic and therapeutic opportunities. In this review, we summarize the current research on the role of ncRNAs in microglial- and astrocyte-mediated neuroinflammation in CNS pathologies.

Alzheimer’s disease

Alzheimer’s disease (AD) is one of the most common age-related neurodegenerative diseases, characterized by memory loss, cognitive impairment, and various neuropsychiatric disorders [24]. The neuropathological hallmarks of AD include the deposition of amyloid beta (Aβ) peptide and intracellular neurofibrillary tangles. Aβ plaques are formed by the cleaved products of the amyloid precursor protein, and neurofibrillary tangles are composed of hyperphosphorylated forms of the microtubule-associated protein tau [25]. Microglia and astrocytes play important roles in mediating neuroinflammation in AD brain tissue. Microglia change their morphology from a ramified to an amoeboid state, and astrocytes show reactive astrogliosis during this disease [26].

Persistent activation of microglia and astrocytes can trigger inflammatory responses, leading to neuronal damage, and ultimately AD [27]. miRNAs play important regulatory roles in glial cell-mediated neuroinflammation in AD. For example, treatment with miR-22-loaded exosomes significantly inhibited microglial activation and inflammatory factor expression by inhibiting pyroptosis [28]. Exosomal delivery of miR-146a suppresses astrocyte inflammation by targeting TRAF6, which promotes synaptogenesis and ameliorates cognitive impairment [29]. In a 3 × Tg AD animal model, miR-155 was significantly upregulated in the mouse brain and Aβ-activated microglia and astrocyte cultures. miR-155 contributes to the production of interleukin 6 (IL-6) and Interferon-β (IFN-β) by sponging off suppressor of cytokine signaling 1 (SOCS1) [30]. It has been reported that cerebrospinal fluid from individuals with AD contains elevated levels of let-7b, and extracellular introduction of let-7b into the cerebrospinal fluid of mice by intrathecal injection resulted in neurodegeneration. Mechanistically, let-7b can activate microglia by acting as a damage-associated molecular pattern against toll-like receptor 7 (TLR7) [31]. In an AD rat model, inhibition of miR-592 attenuated oxidative stress injury in astrocytes by upregulating KIAA0319, thereby alleviating neuronal damage [32]. In another study, miR-135 blockade in astrocytes inhibited neuronal apoptosis and promoted neurite outgrowth by targeting thrombospondin 1 [33]. Aβ aggregation is a major pathogenic factor in AD, and emerging evidence suggests that miRNAs may modulate Aβ production. It was reported that miR-206 expression was significantly upregulated in AD patients, and miR-206 induced inflammation and Aβ release in microglia by targeting insulin-like growth factor 1 (IGF1) [34]. Triggering receptor expressed on myeloid cell 2 (TREM2) is an immunoreceptor primarily found on microglia in the CNS and is critical for Aβ42 peptide clearance. However, miR-34a inhibits TREM2 expression and attenuates the ability of microglia to clear self-aggregating Aβ42 peptides [35]. Moreover, targeted modulation of miR-155 expression regulates the ability of microglia to catabolize fibrillar Aβ1-42; overexpression of miR-155 decreases fibrillar Aβ1-42 catabolism, whereas deletion of miR-155 promotes fibrillar Aβ1-42 catabolism [36]. In addition, the expression of miR-331-3p and miR-9-5p increased in a late-stage AD mouse model. Overexpression of miR-331-3p and miR-9-5p impairs autophagic activity and promotes Aβ formation, whereas their inhibition reduces microglial activation, enhances Aβ clearance, and improves cognition via autophagy activation. Mechanistically, miR-331-3p targets the autophagy receptor sequestosome 1, and miR-9-5p targets the autophagy receptor optineurin [37]. Another study found that miR-138 upregulation in vivo caused an increase in endogenous Aβ42 production as well as changes in inflammatory markers by targeting Sirt1 [38].

In addition to miRNAs, lncRNAs and circRNAs play important regulatory roles in glial cell-mediated neuroinflammation in AD. For example, the expression of lncRNA maternally expressed gene 3 (lncMEG3) decreased in the tissues of AD rats, whereas overexpression of lncMEG3 inhibited astrocyte activation, suppressed inflammatory injury, and alleviated neuronal damage [39]. Neural stem cells (NSCs) can self-renew and generate glial and neuronal lineages. lncRNA urothelial carcinoma-associated 1 (lncUCA1) is upregulated in NSCs and can modulate NSC differentiation. Inhibition of lncUCA1 promotes NSC differentiation into neurons and suppresses its differentiation into astrocytes. Mechanistically, lncUCA1 binds to miR-1 and regulates NSC proliferation and differentiation by targeting Hes1 [40]. In a senescent cell aging model, 7376 circRNAs were identified in primary astrocytes cultured in d-galactose. Among them, CircNF1-419 was significantly upregulated and involved in the modulation of astrocyte autophagy. It enhanced autophagy by binding to dynamin-1 and adaptor protein 2 B1 (AP2B1) [41] (Fig. 1, Table 1).

Fig. 1
figure 1

Brief summary of miRNAs, lncRNAs and circRNAs as a factor controlling neuroinflammation

Table 1 ncRNAs in the regulation of astrocyte- and microglia-mediated neuroinflammation in chronic neurodegenerative diseases

Parkinson’s disease

Parkinson’s disease (PD) is the second most common neurodegenerative disease after AD, and its symptoms include resting tremors, bradykinesia, rigidity, and postural instability. The main pathological features of PD are the loss of dopaminergic neurons and intracellular inclusions containing a-synuclein, called Lewy bodies and Lewy neurites [26]. PD was originally defined as a disease characterized by the loss of dopaminergic neurons; however, an increasing number of studies have shown that PD has an inflammatory component [42,43,44]. Microglia can phagocytose extracellular a-synuclein, and aggregated a-synuclein promotes microglial activation and induces inflammatory responses, which play essential roles in the pathogenesis of PD [45]. In addition, inflammatory mediators secreted by astrocytes and microglia, such as proinflammatory cytokines, reactive oxygen species (ROS), and nitric oxide (NO), regulate the progression of neuronal cell death in PD [46].

Microglial activation leads to exacerbated inflammatory responses in PD, and miRNAs are important regulators of this process [47]. miR-155 was significantly upregulated in a PD mouse model. Its knockout attenuates α-synuclein-induced microgliosis and neurodegeneration, whereas treatment of miR-155 knockout microglia with miR-155 mimics restores the proinflammatory response [48]. In another PD mouse model induced by 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP), miR-7116-5p was downregulated in microglia. Overexpression of miR-7116-5p suppressed microglial activation by targeting tumor necrosis factor α (TNF-α) [49]. Similarly, the expression of miR-335 decreased in a PD model, and miR-335 attenuated proinflammatory responses in microglia by inhibiting leucine-rich repeat kinase 2 (LRRK2) [50]. miR-124 attenuates microglial activation in MPTP-induced PD models. miR-124 inhibits the NF-κB signaling pathway by depleting MEKK3 [51], targets p38 and p62, and promotes autophagy during the inflammatory pathogenesis of PD [52]. Autophagy is closely associated with microglial activation, which attenuates microglial activation and suppresses proinflammatory responses. TREM2 inhibits autophagy via mTOR signaling. However, TREM2 was identified as a target of miR-3473b in an MPTP-induced PD model. miR-3473b antagomir treatment inhibited microglial activation and promoted autophagy [53]. Furthermore, the phosphatase and tensin homolog (PTEN), an inducer of autophagy, is targeted by miR-19a-3p in microglia. Exosome overloading with miR-19a-3p suppresses autophagy in recipient microglia by targeting the PTEN/AKT/mTOR pathway [54].

The NOD-like receptor protein 3 (NLRP3) inflammasome is highly expressed in microglia and has been implicated in PD pathogenesis. Importantly, miR-7 targets NLRP3 expression and inhibits NLRP3 inflammasome activation. The injection of miR-7 mimics into PD model mice attenuated microglial activation and dopaminergic neuron degeneration [55]. Furthermore, lncRNA small nucleolar RNA host gene 1 (lncSNHG1) functions as a competing endogenous RNA (ceRNA) for miR-7 to modulate NLRP3 expression, which promotes microglial activation and the NLRP3 inflammasome [56]. In addition, lncRNA growth arrest-specific 5 (lncGAS5) modulates NLRP3 expression in both in vivo and in vitro models of PD and upregulates NLRP3 via competing miR-223-3p [57]. In addition to functioning as ceRNAs, lncRNAs regulate inflammasomes by binding to proteins. In an MPTP-induced PD model, MALAT1 recruits the enhancer of zeste homolog 2 (EZH2) to the nuclear factor-like-2 (NRF2) promoter and inhibits NRF2 expression, thereby facilitating inflammasome activation in PD microglia and mouse models [58]. In addition, lncRNAs regulate neuroinflammation by modulating NF-κB signaling in PD. Follistatin-like 1 (FSTL1) is an activator of NF-κB signaling pathways, and reduction of lncHOXA11-AS inhibits FSTL1 expression by upregulating miR-124-3p, thereby attenuating microglial activation and neurological damage in PD models [59]. PKC-δ, which can regulate NF-κB activation, is upregulated by lincRNA-p21 by sponging miR-181. Interestingly, p53, lincRNA-p21, miR-181, and PKC-delta form a double-negative feedback loop that promotes persistent microglial activation in PD [60] (Fig. 1, Table 1).

Amyotrophic lateral sclerosis

Amyotrophic lateral sclerosis (ALS) is a fatal adult-onset neurodegenerative disease that affects motor neurons (MN) in the brainstem, spinal cord, and motor cortex [61]. The clinical features include muscle stiffness and twitching, limb weakness, cognitive impairment, and hyperreflexia. Once considered a motor neuron disease, ALS is now recognized as a multisystem, multicellular disorder [62]. Neuroinflammation is readily observed in imaging studies of human patients with ALS and rodent ALS models and is characterized by microgliosis and astrocytosis. Activated microglia and astrocytes can further damage the MN and contribute to neuronal death by secreting proinflammatory cytokines and apoptosis-inducing molecules [26].

Mutations in superoxide dismutase 1 (SOD1) are involved in the pathogenesis of ALS. The SOD1 mouse, which expresses human SOD1 protein with ALS mutations, can recapitulate the disease and is the most widely used model for ALS [63]. A recent study showed alterations in neuroprotective and neuroinflammatory effects in the spinal cord of SOD1 mice. In the symptomatic stage, glial function is significantly depressed, cell-to-cell communication is reduced, and miR-155 is upregulated. In contrast, astrocytosis, microgliosis, and neuroinflammation are evident during the symptomatic stages. Moreover, several inflammation-related miRNAs, including miR-155, miR-146a, miR-125b, miR-124, and miR-21, were significantly upregulated, suggesting that these miRNAs may play essential roles in modulating neuroinflammation in ALS [64]. Indeed, miRNAs are widely involved in the regulation of microglial- and astrocyte-mediated inflammatory responses in ALS. In a SOD1 mouse model, miR-155 knockout reversed the abnormal molecular signature and phagocytic function of microglia. In addition, treatment of SOD1 mice with anti-miR-155 delayed disease onset and prolonged survival [65]. The anti-inflammatory protein A20 is induced in microglia upon inflammatory BzATP stimulation, and A20 is regulated by miR-125b. Inhibition of miR-125b suppresses NF-κB activation and the release of toxic factors, thereby protecting MN from death induced by activated microglia [66]. The expression of miR-146a is decreased in astrocytes isolated from the cortex of symptomatic SOD1 (mSOD) mice, whereas its target genes IRAK1 and TRAF6 are upregulated. Overexpression of pre-miR-146a in mSOD astrocytes attenuates the aberrant phenotype of astrocytes, including the restoration of GFAP, S100B, vimentin, Cx43, and HMGB1. In contrast, the transfection of anti-miR-146a into wild-type astrocytes reproduces the aberrant phenotype of mSOD1 astrocytes [67]. miRNAs also regulate the differentiation of neurons and astrocytes. miR-124 is downregulated in the spinal cord and brainstem of SOD1 mice and is associated with the astrocytic differentiation of NSCs by targeting Sox2 and Sox9 [68].

Glial cells normally support neurons; however, in ALS, the homeostatic role of glial cells is often lost and replaced by deleterious effects on neurons. Glial cells may respond to the dying neurons and exacerbate their degeneration [69]. Therefore, neuron–glial crosstalk is of great importance in ALS, and miRNAs are critical regulators of this process. A recent study showed that astrocytes from patients with ALS are toxic to MN and that this toxicity is mainly mediated by astrocyte-derived extracellular vesicles (ADEV). Moreover, miR-494-3p was significantly downregulated in ADEV and regulated its target gene SEMA3A in MN. SEMA3A contributes to reduced neurite growth and MN death in ALS. Therefore, reduced miR-494-3p expression in ADEV leads to MN loss and ALS pathogenesis [70]. In addition, MN-derived miRNAs directly modulate glial cell phenotypes. MN-specific miR-218 is highly enriched in MN and is released extracellularly in ALS rat models. Astrocytes that do not express miR-218 take up extracellular miR-218 released from dying MN, and miR-218 downregulates the glutamate transporter EAAT2 in astrocytes, thus contributing to astrogliosis and progressive neuronal damage [71] (Fig. 1, Table 1).

Multiple sclerosis

Multiple sclerosis (MS) is a chronic neurodegenerative disease of the central nervous system (CNS) that occurs primarily in young adults [72]. It is mainly considered an autoimmune disease characterized by inflammation and demyelination of neurons. The progressive pathological processes of MS include BBB breakdown, multifocal inflammation, reactive microgliosis, astrocytosis, demyelination, and axonal degeneration [73]. Increasing evidence suggests that MS progression and symptoms are closely correlated with the maintenance of persistent, low-grade inflammation driven by microglia and astrocytes [74].miRNAs are important modulators of microglial- and astrocyte-mediated inflammatory responses in MS. miRNA profiling shows that 20 miRNAs, including miR-155, miR-146a, miR-223, miR-142, miR-34a, and miR-326, are highly upregulated in active MS lesions, among which miR-155 regulates neuroinflammation in MS [75]. miR-155 promotes proinflammatory responses in microglia and astrocytes by targeting SOCS1 [76, 77]. In addition to its direct proinflammatory effect on glial cells, miR-155 can exacerbate neuroinflammation in MS by targeting focal adhesions and tight junctions, resulting in increased BBB permeability and peripheral leukocyte infiltration into the CNS [78]. In contrast, miR-146a has a heterogeneous role in MS. In an experimental autoimmune encephalomyelitis (EAE) mouse model, administration of miR-146a mimics facilitated M2 microglial polarization and promoted oligodendrocyte progenitor cell (OPC) differentiation and remyelination, suggesting the therapeutic functions of miR-146a in MS [79]. Furthermore, inflammatory microglial transfer of miR-146a to neurons via extracellular vesicles (EVs) leads to a significant decrease in the density of dendritic spines and excitatory synapses, and sustained exposure to miR-146a-enriched EVs results in pathological synapse loss and synaptic dysfunction [80]. The role of miR-223 in inflammatory responses is also controversial in MS. In an EAE model, miR-223 knockout ameliorated pathogenic CNS inflammation, demyelination, and clinical symptoms of EAE [81]. Mechanistically, miR-223 inhibits autophagy by targeting ATG16L1; therefore, miR-223 deficiency promotes microglial autophagy and increases the number of resting microglia. Moreover, miR-223 deficiency leads to delayed onset of EAE, but the disease severity differed [82]. miR-223 deficiency has little effect on the proinflammatory phenotype of microglia and macrophages; however, miR-223 is essential for M2 polarization and phagocytosis of microglia and macrophages, and miR-223 knockout impairs CNS remyelination and myelin debris clearance. miR-124 and miR-30a also regulate the inflammatory response of microglia to EAE. In vivo administration of miR-124 reduces EAE progression by promoting microglial quiescence and deactivating macrophages by targeting the C/EBP-α-PU.1 pathway [83]. Conversely, miR-30a contributes to the microglial inflammatory response, and transplantation of miR-30a-modified microglia exacerbates EAE progression [84].

The inflammatory cytokine interleukin-17 (IL-17) is a key regulator in autoimmune diseases. Increasing evidence indicates that miRNAs are broadly involved in the IL-17-mediated proinflammatory response in MS [85]. miR-23b is one of the prominent miRNAs involved in IL-17-associated autoimmune inflammation in MS. IL-17 decreases the expression of miR-23b in astrocytes, thereby inhibiting the expression of inflammatory cytokines by targeting TAB2, TAB3, and IKK-α. Therefore, IL-17 exacerbates MS progression by attenuating miR-23b expression and increasing proinflammatory cytokine expression [86]. In an EAE model, IL-17 increases miR-873 level in astrocytes and mouse brain tissue, which enhances the expression of inflammatory cytokines and exacerbates MS progression via regulation of the A20/NF-κB pathway [87]. Similarly, miR-409-3p and miR-1896 induced by IL-17 stimulation coordinately promote proinflammatory responses in reactive astrocytes by modulating the SOCS3/STAT3 pathway [88].

Emerging evidence suggests that IL-9 plays a regulatory role in autoimmune responses in MS. A recent study characterized an IL-9-triggered lncRNA, Gm13568, in astrocytes, which promotes proinflammatory responses in active astrocytes and contributes to EAE pathogenesis. Gm13568 interacts with CBP/P300 and stimulates Notch1 pathway activation, thereby increasing the production of inflammatory cytokines [89]. lncRNAs also modulate microglial polarization in MS. Microarray screening shows that 120 lncRNAs are differentially expressed in M2-polarized microglia versus resting microglia, among which the lncGAS5 is significantly downregulated. lncGAS5 inhibits TRF4 transcription by recruiting PRC2, thereby attenuating M2 microglial polarization. Knockdown of lncGAS5 in transplanted microglia ameliorates EAE progression and facilitates remyelination [90]. Sulfasalazine (SF) is an anti-inflammatory drug that improves outcomes in patients with MS. A recent study has shown that SF suppresses M1 microglial polarization and promotes remyelination. Mechanistically, SF blocked AKT2-NF-κB signaling through the effects of lncHOTAIR and miR-136-5p in microglia [91] (Fig. 1, Table 1).

Ischemic stroke

Ischemic stroke (IS) is a leading cause of death and disability worldwide, resulting from the occlusion of a cerebral artery that interrupts cerebral blood flow and causes rapid loss of brain function [92]. Reperfusion therapy is an effective therapeutic approach for reducing primary injury; however, it can also induce cerebral ischemia–reperfusion (I/R) injury, which causes secondary neuronal damage and death [93]. Post-IS inflammation mediated by microglia and astrocytes plays a dual role in brain tissue damage and repair [94]. Microglia are rapidly activated after IS and produce proinflammatory cytokines that cause tissue injury. In contrast, M2-phenotype microglia release anti-inflammatory cytokines that contribute to functional recovery after IS. Reactive astrogliosis may exacerbate ischemic lesions and hinder axonal regeneration. However, it also contributes to neuroprotection and neurological recovery [95].miRNAs are involved in the regulation of various pathogenic mechanisms underlying tissue injury after stroke, including inflammatory responses, excitotoxicity, oxidative stress, mitochondrial dysfunction, and BBB dysfunction [96]. Importantly, miRNAs are involved in glia-mediated inflammation by modulating the activation, polarization, proliferation, and apoptosis of microglia and astrocytes. Several miRNAs have been shown to ameliorate glia-mediated inflammatory responses [97]. miR-146a is significantly upregulated in microglia after oxygen–glucose deprivation (OGD) [98], and miR-146a-loaded exosomes can inhibit microglial activation and neuroinflammation via the IRAK1/TRAF6 pathway [93]. Similarly, miR-1906 and miR-669c can ameliorate post-IS neuroinflammation by modulating the TLR signaling pathway; miR-1906 directly targets TLR4, whereas miR-669c inhibits the canonical adaptor protein MyD88, thereby blocking TLR4 signaling and reducing inflammatory responses [99, 100]. miR-124 contributes to neuroprotection and functional recovery and may serve as a promising candidate for treating IS [101]. Recent studies further support the beneficial role of miR-124 in IS; miR-124 transported by M2 microglia-derived exosomes can suppress neuronal apoptosis and inhibit glial scar formation, thereby improving the outcome of IS [102, 103]. Similarly, miR-424 attenuates brain injury and promote functional recovery after IS, and its effects are mediated by attenuating microglial activation and astrogliosis [104, 105]. In addition, overexpression of let-7c-5p inhibits microglial activation and attenuates brain damage after IS by targeting caspase 3 [106]. Microglial phagocytosis of stressed but viable neurons is considered detrimental to the brain, whereas EV-derived miR-98 can prevent microglial phagocytosis of salvageable neurons by targeting PAFR, thereby reducing neuronal death during IS [107]. Moreover, miR-98 protects the BBB from proinflammatory monocyte infiltration, thereby preventing further microglial activation and improving neurological outcomes after IS [108].

However, miRNAs can also aggravate IS-induced brain injury. The miR-15a/16-1 cluster has opposite regulatory roles to miR-98 in BBB permeability; knockout of the miR-15a/16-1 cluster attenuates peripheral immune cell infiltration and inhibits M1 microglia [109]. In a mouse model of middle cerebral artery occlusion (MCAO), miR-3473b levels increased in the cortex and striatum. Upregulation of miR-3473b exacerbates the pathogenesis of IS by promoting microglia-mediated neuroinflammatory injury, and the associated mechanism involves the regulation of SOCS3 [110]. In another MCAO model, miR-181c aggravated brain ischemia–reperfusion injury by promoting microglial and neuronal apoptosis by regulating apoptosis-related genes, including BCL-2 and BAX [111]. miR-148b and miR-365 modulate neurogenesis in IS; miR-148b inhibits the proliferation and differentiation of NSC into neurons and astrocytes by regulating Wnt/β-catenin signaling, thereby attenuating the recovery of neurological function after stroke [112]. miR-365 suppresses the conversion of astrocytes into mature neurons and subsequently exacerbates ischemic injury by targeting PAX6 [113].lncRNAs and circRNAs have recently emerged as critical modulators of IS. The dysregulation of lncRNAs is involved in regulating microglial inflammation by its effects on microglial activation and polarization [114, 115]. In a mouse model of MCAO, overexpression of lnc1810034E14Rik inhibited microglial activation by reducing p65 phosphorylation, thereby alleviating brain damage [116]. lncRNAs NEAT1, H19, TUG1, and GAS5 can shift microglial polarization in IS, and lncNEAT1 attenuates M1 microglial polarization to suppress OGD/R-induced injury [117]. In contrast, lncH19 promotes M1 microglial polarization and contributes to neuroinflammation [118]. Similar to H19, lncTUG1 can facilitate microglial polarization toward the M1 phenotype, and this process is mediated by miR-145a-5p [119]. lncGAS5 increases M1 microglial polarization and decreases M2 microglial polarization by upregulating Notch1 expression via miR-146a [120]. In addition, lncRNAs are involved in glial cell apoptosis. Downregulation of MALAT1 inhibits astrocyte apoptosis and protects against cerebral I/R injury; the associated mechanism is related to miR-145 and its target AQP4 [121]. In contrast, lncNespas knockdown exacerbates the I/R-induced microglial apoptosis and inflammatory responses. lncNespas can block the interaction between TRIM8 and TAK1 and attenuate the K63-linked polyubiquitination of TAK1, resulting in the inactivation of TAK1 and NF-κB signaling [122]. Moreover, microglial activation-induced lncRNA U90926 promotes neutrophil infiltration via a mechanism involving the regulation of CXCL2. U90926 interacts with MDH2 and prevents the binding of MDH2 to the CXCL2 3ʹ UTR, thereby inhibiting MDH2-mediated decay of CXCL2 mRNA [123]. lncRNAs also serve as therapeutic targets for IS by regulating NSC differentiation. Furthermore, inhibiting lncPnky in NSCs promotes their differentiation into neurons and astrocytes, leading to improved functional recovery after IS [124]. In addition, circRNAs are involved in IS-associated autophagy. In a mouse model of MCAO, circHECTD1 was highly expressed and significantly upregulated in the ischemic brain tissue. circHECTD1 promotes autophagy in astrocytes by targeting the miR-142/TIPARP axis [125]. circSHOC2 is highly expressed in astrocyte-derived exosomes. It functions as an endogenous miR-7670-3p sponge to regulate SIRT1 expression, resulting in reduced neuronal autophagy and ameliorating neuronal damage after IS [126] (Fig. 1, Table 2).

Table 2 ncRNAs in the regulation of astrocyte- and microglia-mediated neuroinflammation in acute neurodegenerative disease

Traumatic brain injury

Traumatic brain injury (TBI) is a leading cause of mortality in developed countries. The pathology of TBI is complex and can be divided into two main stages: primary and secondary injuries. Primary injury occurs during lesion formation, resulting in contusion and hemorrhage. Secondary injury is caused by the complicated processes of initial impact and is characterized by various neuropathological processes, including ischemia, oxidative stress, excitotoxicity, apoptosis, necrosis, and neuroinflammation [127]. Among these, neuroinflammation is an important pathological process, and microglia and astrocytes are considered critical players in initiating inflammatory responses and determining the extent of damage during TBI [128].

In a TBI rat model, the expression of miR-155 and miR-142 was highly upregulated in the perilesional cortex; miR-155 was predominantly expressed by activated astrocytes, whereas miR-142 expression was associated with microglia, macrophages, and lymphocytes. miR-155 and miR-142 contribute to astrocyte activation and brain inflammation after TBI [129]. In particular, miR-155 plays a critical role in progressive neuroinflammatory responses in TBI. Microglia-derived microparticles loaded with proinflammatory mediators, including miR-155, promote microglial activation and persistent neuroinflammatory responses [130]. Moreover, miR-155 inhibition suppresses post-traumatic neuroinflammatory responses and improves neurological recovery after TBI in mice [131]. miR-200b modulates microglial inflammatory processes in TBI. The expression of miR-200b is decreased in activated microglia, and the downregulation of miR-200b leads to increased inflammatory responses via cJun/MAPK signaling [132]. In addition, miRNAs modulate astrocyte differentiation and proliferation in TBI. Activated astrocytes can regulate NSC differentiation by secreting various cytokines, such as leukemia inhibitory factor (LIF) and ciliary neurotrophic factor (CNTF), which promote the premature generation of astrocytes via the activation of JAK and STAT signaling pathway. However, several proteins involved in JAK2/STAT3 signaling are direct targets of the miR-17-92 cluster. Therefore, the miR-17-92 cluster suppresses astrocytogenesis and increases neurogenesis, improving motor coordination as observed in brain-injured mice [133]. Furthermore, miR-9-5p is significantly upregulated in brain tissue after TBI. Inhibition of miR-9-5p upregulates the expression of its target gene, Thbs-2, in astrocytes, thereby promoting the proliferation of astrocytes and the release of astrocyte-derived neurotrophic factors, leading to the recovery of neurological function [134].

Exosomes are associated with the cell-to-cell crosstalk involved in immune regulation in TBI, and several exosome-carried miRNAs have been shown to modulate TBI neuroinflammation [135]. In TBI, miRNA-loaded exosomes are involved in communication between glia and neurons. In a model of repetitive mild TBI, the expression of miR-124-3p is significantly upregulated in microglial exosomes. Treatments with miR-124-3p-loaded exosomes promote anti-inflammatory M2 polarization in microglia and alleviate neurodegeneration by targeting the Rela/ApoE signaling pathway [136]. A similar study involving TBI shows that Exo-miR-124-3p contributes to M2 microglial polarization and attenuates neuronal inflammation via transfer into neurons; these effects are mediated by PDE4B/mTOR signaling [137]. Neuron-derived exosomes regulate glial inflammatory responses in TBI. For example, neuron-derived exosomes loaded with miR-21-5p are phagocytosed by microglia and promote M1 microglial polarization [138]. In addition, miRNA-loaded exosomes are involved in the cellular communication between astrocytes and microglia. Moreover, activated astrocyte-derived exosomes are enriched with miR-873a-5p, which can suppress microglial inflammatory responses by promoting M2 microglial polarization after TBI [139].

Several lncRNAs are implicated in the onset and progression of TBI. lncRNAs modulate microglial inflammation in TBI [140]. For example, lncMEG3 facilitates Nlrp3-mediated microglial activation and inflammatory responses by targeting miR-7a-5p [141]. The lncRNA KCNQ1OT1 serves as an endogenous miRNA sponge for miR-873-5p, thereby promoting TRAF6 expression and ultimately contributing to M1 microglial polarization [142]. HOTAIR is highly expressed in activated microglia, where it binds to the MYD88 protein and increases its stability by inhibiting Nrdp1-mediated ubiquitination of MYD88. Downregulation of HOTAIR suppresses microglial activation and the release of inflammatory factors [143]. In addition, lncRNAs are involved in regulating astrocytic inflammation in TBI. lncGm4419 level increases significantly in astrocytes after TBI and can enhance TNF-α expression by competitively binding to miR-466l to promote astrocyte apoptosis [144]. Cerebral edema, an important neurological complication of TBI, is characterized by early astrocyte swelling. MALAT1 reduces astrocyte swelling and improve brain edema; its effects are mediated by reducing AQP4 and IL-6 expression [145]. In addition, astrocyte-derived EVs carrying the lncRNA NKILA contribute to the recovery of injured neurons after TBI. Mechanistically, NKILA promoted NLRX1 expression by competitively binding to miR-195 [146] (Fig. 1, Table 2).

Spinal cord injury

Spinal cord injury (SCI) is a major cause of morbidity and mortality worldwide, resulting in long-term cognitive or motor dysfunction [147]. Similar to TBI, mechanical insults induce primary damage to the spinal cord, and a series of pathological cascades cause secondary damage, leading to neuronal and glial cell death, inflammation, and ischemia [148].

Astrogliosis is characterized by astrocyte proliferation and hypertrophy, resulting in the formation of glial scars that act as physical barriers to axonal regeneration. However, in the early stages of SCI, astrogliosis is essential for preventing advanced injury [149]. Various miRNAs have been shown to modulate the astrocyte phenotype and function in SCI. The RNase III ribonuclease Dicer1 is required for the formation of mature miRNAs, and deletion of Dicer1 prevents astrocyte proliferation after SCI. However, treatment with the miR-17-5p mimic corrects the defective proliferation of Dicer1 knockout astrocytes. Furthermore, the suppression of miR-17-5p can prevent lipopolysaccharide-induced astrocyte proliferation [150]. The levels of astrocyte-enriched miR-145 decrease after SCI, and the downregulation of miR-145 in astrocytes promotes astrocyte growth, motility, and hypertrophy by enhancing GFAP and c-myc expression, which is beneficial for spinal cord tissue repair [151]. In addition, miR-21 expression is significantly upregulated in response to SCI. Overexpression of miR-21 in astrocytes contributes to astrocyte secretion and proliferation and promotes recovery after SCI; these effects are mediated by PTEN/ PI3K/Akt/mTOR signaling [152].

Activated microglia promote secondary tissue damage in SCI. Similarly, miRNAs have been implicated in the modulation of microglial activation and polarization in SCI [153]. miR-124 is a well-studied miRNA that regulates microglial response after SCI. In a rat SCI model, the delivery of miR-124 to rat microglia inhibited microglial activation and subsequently attenuated neuroinflammation, suggesting that miR-124 is a therapeutic target for suppressing inflammation in SCI [154]. In addition, neuron-derived exosomes loaded with miR-124 facilitated functional behavioral recovery in SCI by inhibiting the activation of microglia and astrocytes [155]. In addition, miR-124 regulates the phagocytic response of microglia. Moreover, docosahexaenoic acid (DHA) exerts neuroprotective effects in SCI, and these effects are mediated by a miR-124-dependent reduction in microglial phagocytic activity [156]. Similarly, exosomes secreted by mesenchymal stem cells (MSCs) facilitate functional behavioral recovery by modulating microglial activation. Exosome-carried miR-216a-5p shifts microglia from the M1 to the M2 phenotype by targeting the TLR4/NF-κB/PI3K/AKT signaling pathways [157]. In addition, miR-340-5p ameliorates neuroinflammation and promotes functional recovery after SCI. Mechanistically, miR-340-5p targets P38 and inhibits the MAPK pathway to ameliorate the SCI-induced inflammatory responses [158].

Interestingly, both a contusive SCI mouse model and RNA-seq of the SCI epicenter in the acute and chronic stages showed that the protein-coding gene Zeb2 and its antisense lncRNA Zeb2os were upregulated and colocalized in astrocytes. Knockdown of Zeb2 and lncZeb2os attenuated reactive astrogliosis, demonstrating similar roles of Zeb2 and lncZeb2os in the astrocytic response after SCI. lncZeb2os can positively regulate Zeb2 expression, and these effects may be mediated by RNA–RNA, RNA–DNA, or RNA–protein interactions [159, 160]. Importantly, several lncRNAs (including lncGm37494, lncMALAT1, lncXIST, lncFtx, lncLEF1-AS1, lncCCAT1, and lncF630028O10Rik) have been confirmed as endogenous miRNA sponges that regulate the inflammatory responses of microglia and astrocytes after SCI [161,162,163,164,165,166,167]. For example, TLR4-induced lncF630028O10Rik promotes microglial pyroptosis after SCI by targeting the miR-1231-5p/Col1a1 axis, thus providing a therapeutic target for attenuating neuroinflammation in SCI [167]. In addition, lncRNAs interact with proteins to regulate SCI. KLF4 is a conserved transcription factor involved in microglial and astrocytic activation in SCI. lncSNHG5 can directly bind to KLF4 and increase its expression, thereby facilitating the viability of microglia and astrocytes [168]. circRNAs have also been implicated in the regulation of SCI progression. A recent circRNA profile in an SCI rat model showed that 360 circRNAs were differentially expressed in the spinal cord tissues, among which circRNA_01477 expression was significantly decreased. Downregulation of circRNA_01477 suppresses astrocyte proliferation and migration [169] (Fig. 1, Table 2).

Viral encephalitis

Although the CNS is not a common viral target, many viruses can infect the brain and induce neuroinflammation, which is characterized by severe neuronal injury and microgliosis/astrogliosis. Microglia and astrocytes actively respond to both RNA and DNA viruses by releasing various inflammatory mediators to activate the immune system and combat infection [170,171,172].

Japanese encephalitis virus (JEV) is a mosquito-borne ssRNA virus that causes acute inflammatory diseases of the CNS called Japanese encephalitis (JE). JE is the most common viral encephalitis in the Asia–Pacific region, killing one-third of patients and leaving nearly half of the survivors with permanent neuropsychiatric sequelae [173]. JEV directly infects microglia and astrocytes and induces glial activation. Accumulating evidence has implicated miRNAs as essential regulators of microglia-mediated neuroinflammation in JEV infections. In response to JEV infection, miR-146a was significantly upregulated in microglia and functioned as a negative feedback regulator in JEV-induced inflammatory responses [174]. miR-146a also facilitated JEV replication by inhibiting NF-κB activity and antiviral JAK–STAT signaling [175]. miR-155 is also upregulated after JEV infection; however, its role remains controversial. One study showed that increased miR-155 expression further exacerbates JEV-induced neuroinflammation by targeting SHIP1 [176]. Another study showed that miR-155 inhibited microglial activation and innate immune responses, thereby suppressing JEV replication [177]. In addition, miR-29b and miR-301a facilitate microglial activation by modulating negative regulators of NF-κB signaling; miR-29b targets TNFAIP3, and miR-301a inhibits NKRF, leading to the activation of NF-κB activity and enhanced JEV-induced inflammatory responses [178, 179]. Let-7a/b enhances TNF-α production in microglia by interacting with TLR7 and NOTCH signaling, and exosomes overloading Let-7a/b can induce neuronal death through caspase activation [180]. miRNAs also modulate astrocyte-mediated neuroinflammation in JEV infection; miR-19b-3p and miR-15b are upregulated in JEV-infected astrocytes, both exacerbating JEV-induced inflammatory responses via the negative regulation of RNF11 and RNF125, respectively [173, 181].

Infection with human immunodeficiency virus type 1 (HIV-1) leads to the progressive weakening of the host's immune system and causes acquired immunodeficiency syndrome (AIDS). HIV-1 infection is often associated with chronic brain inflammation, and approximately 70% of patients develop HIV-1-associated neurological disorders (HAND) [182]. HIV can infect the resident cells of the CNS, including microglia and astrocytes, resulting in severe neuroinflammatory consequences [183]. HIV-1 transactivator of transcription (Tat) is an HIV-1-encoded viral protein that has attracted considerable attention because of its toxicity to CNS cells. HIV-1 Tat influences the functional dynamics of microglia and that miRNAs are important regulators of this process. Exposure of microglia to Tat leads to increased miR-34a expression, which targets the NF-κB negative regulator, NLRC5, and promotes microglial activation [184]. Tat decreases miR-124 expression by mediating DNA methylation of the miR-124 promoter, resulting in the upregulation of STAT3 and subsequent microglial activation [185]. miR-146a levels increase in HIV-1-infected primary human fetal microglia, and MCP-2 has been identified as a target of miR-146a [186]. miRNAs can modulate oxidative stress in microglia upon Tat treatment; exposure of microglia to Tat downregulates the level of miR-17, which in turn increases the expression of NOX2 and NOX4, thereby promoting ROS generation [182]. Tat upregulates the expression of miR-505 in microglia. miR-505 contributes to the production of mitochondrial superoxide and microglial senescence-like phenotypes by negatively regulating SIRT3 expression [187]. In addition, HIV-1-latently infected astrocytes can express and secrete Tat, which induces pathological and neurobehavioral changes in the CNS. In addition, Tat can cause astrocyte-mediated neuronal neurotoxicity via the miR-320a/VDAC1 axis, and overexpression of miR-320a inhibits Tat-mediated ATP release and prevents neuronal death [188]. miR-132 expression is induced by Tat through CREB phosphorylation in astrocytes, and the uptake of exosome-delivered miR-132 into neurons induces the downregulation of BDNF and MECP2, which impairs neurite outgrowth and neuronal survival [189]. Similarly, miR-29b is present in astrocyte-derived exosomes and is taken up by neurons, leading to attenuated PDGF-B expression and subsequent neuronal dysfunction [190]. In addition, miR-155 and miR-181a have been implicated in HIV-1 replication. The overexpression of these two miRNAs promotes HIV-1 replication in astrocytes by modulating SAMHD1 [191]. In response to HIV-1 infection, lncHEAL expression is upregulated in microglia, macrophages, and T lymphocytes, facilitating HIV-1 replication by interacting with the RNA-binding protein, FUS [192]. In addition, lncBACE1-AS has been implicated in Tat-mediated astrocytic amyloidosis. It binds to HIF-1α, forming the BACE1/lncBACE1-AS RNA duplex and increasing BACE1 protein and astrocytic amyloidosis [193] (Fig. 1, Table 3)

Table 3 ncRNAs in the regulation of astrocyte- and microglia-mediated neuroinflammation in CNS infectious diseases

.

Bacterial encephalitis

Bacterial infection of the CNS is a significant global public health problem with high mortality and morbidity rates [194]. Based on the affected anatomical regions, CNS infections can be classified as encephalitis, meningitis, or myelitis, with bacterial meningitis being the most common and severe disease [195]. Escherichia coli (E. coli), Streptococcus pneumoniae (S. pneumoniae), Listeria monocytogenes (L. mo), Mycobacterium tuberculosis (M. tb), Neisseria meningitidis, and Haemophilus influenzae type b are the major causes of meningitis [196, 197]. Bacteria multiply in the subarachnoid space and release compounds related to pathogen-associated molecular patterns (PAMPs), such as lipopolysaccharides, peptidoglycan, flagellin, DNA, and lipoteichoic acid. Microglia and astrocytes can induce immune responses by recognizing PAMPs through pattern recognition receptors, thereby facilitating the elimination of invasive bacteria [198]. However, an excessive inflammatory response can cause neuronal damage and death. Therefore, the intensity and duration of inflammatory responses should be maintained at appropriate levels, and miRNAs are critical regulators of glia-mediated neuroinflammation during bacterial infections [199].

Our group has previously profiled the expression of miRNAs in E. coli-infected astrocytes [200]. Transcriptome data showed that 16 miRNAs were upregulated and 11 miRNAs were downregulated compared to the control group. We further characterized the roles of the most significantly upregulated miRNAs, miR-155 and miR-146a [201]. These two miRNAs collectively attenuate E. coli-induced neuroinflammatory responses through negative feedback regulation of the TLR-mediated NF-κB and EGFR/NF-κB signaling pathways. Lipopolysaccharide (LPS), a major component of several Gram-negative bacteria, is known to be an important endogenous promoter of neuroinflammation. Studies suggest that low expression of miR-138-5p after LPS administration may contribute to the activation of NLRP3/caspase-1 in microglia, leading to hippocampal neuroinflammation [202, 203]. In addition, miR-155 modulated brain inflammation via multiple mechanisms during L. mo infection. Peripheral miR-155 contributes to inflammation by promoting the recruitment of T lymphocytes, whereas microglial miR-155 plays a dual role: miR-155 can enhance M1 microglial polarization and suppress inflammatory responses by targeting TAB2 [204]. miR-30c-5p is beneficial during Cryptococcus neoformans infection as it attenuates the release of inflammatory cytokines and promotes microglial survival [205]. Since M. tb is internalized and replicates within microglia, miR-124 is thought to protect against M. tb infection by promoting microglial apoptosis and facilitating M. tb clearance. Mechanistically, miR-124 directly targets STAT3 and abrogates the anti-apoptotic effects of STAT3 signaling [206]. In an S. pneumoniae-induced bacterial meningitis model, miR-141-3p expression decreased in the brain tissue and astrocytes. Overexpression of miR-141-3p suppresses astrocyte activation and inflammatory responses by negatively regulating HMGB1 [207]. miR-135a is also downregulated in brain tissue of S. pneumoniae-infected rats, facilitating proliferation and inhibiting apoptosis of astrocytes by targeting HIF-1α [208] (Fig. 1, Table 3).

Conclusion and future directions

Astrocytes and microglia, the most important components of the innate immune system in the CNS, constantly monitor the brain microenvironment under normal conditions. They can be activated following inflammation, infection, or trauma and play an important role in the pathological process and development of CNS diseases. The roles of miRNAs, lncRNAs, and circRNAs in regulating astrocyte- and microglia-mediated neuroinflammation during CNS diseases are summarized in this review. It is well-known that ncRNAs regulate the expression of genes involved in various signaling pathways in astrocytes and microglia that contribute to neuroinflammation. Disruption of these signaling cascades can be achieved by the overexpression or inhibition of ncRNAs, allowing astrocytes and microglia to revert to the neuroprotective phenotype involved in CNS repair and recovery.

Different types of brain cells work together to maintain healthy brain activity. Under normal physiological conditions, neurons and glia communicate with and integrate signals from the surrounding cells and the environment to maintain CNS stability. Disturbances in the neuron–glia crosstalk contribute to various pathological states of CNS diseases. Under pathological conditions, the release, recognition, and uptake of molecules or other cellular components by neurons, astrocytes, and microglia are altered. Therefore, a better understanding of the crosstalk between neurons, astrocytes, and microglia is essential to understand the mechanisms of neuroinflammation in CNS diseases. This review summarizes the ncRNAs involved in intercellular communication in CNS diseases (Fig. 2). In particular, ncRNA research has opened new therapeutic strategies for the treatment of neuroinflammation, because ncRNAs can be packaged as exosomes, which are better able to cross the BBB [209]. For instance, in AD, treatment with miR-22-loaded exosomes significantly inhibited M1 microglial differentiation and the expression of inflammatory factors [28]. In IS, miR-146a-loaded exosomes inhibit microglial activation and neuroinflammation via the IRAK1/TRAF6 pathway [93]. Therefore, the development of therapies using ncRNA cocktails and exosome gene therapy has great potential and opens new avenues for personalized treatment of neurological diseases.

Fig. 2
figure 2

Brief summary of ncRNAs crosstalking between neurons, astrocytes, and microglia in CNS diseases

Availability of data and materials

Not applicable.

Abbreviations

CNS:

Central nervous system

BBB:

Blood–brain barrier

GFAP:

Glial fibrillary acidic protein

ncRNAs:

Non-coding RNAs

miRNA:

MicroRNA

lncRNA:

Long non-coding RNA

circRNA:

Circular RNA

UTRs:

Untranslated regions

nt:

Nucleotides

mRNA:

Messenger RNA

AD:

Alzheimer’s disease

TLR7:

Toll-like receptor 7

Aβ:

Amyloid β

IGF1:

Insulin-like growth factor 1

TREM2:

Triggering receptor expressed on myeloid cells-2

MEG3:

Maternally expressed gene 3

NSC:

Neural stem cells

UCA1:

Urothelial carcinoma-associated 1

AP2B1:

Adaptor protein 2 B1

PD:

Parkinson’s disease

MPTP:

1-Methyl-4-phenyl-1,2,3,6-tetra hydropyridine

PTEN:

Phosphatase and tensin homolog

NLRP3:

Nod-like receptor protein 3

SNHG1:

Small nucleolar RNA host gene 1

ceRNA:

Competing endogenous RNA

EZH2:

Enhancer of zeste homologue 2

NRF2:

Nuclear factor-like-2 factor

FSTL1:

Follistatin-like 1

ALS:

Amyotrophic lateral sclerosis

MN:

Motor neurons

SOD1:

Superoxide dismutase 1

mSOD:

Symptomatic SOD1

ADEV:

Astrocyte-derived extracellular vesicles

MS:

Multiple sclerosis

EAE:

Experimental autoimmune encephalomyelitis

OPC:

Oligodendrocyte progenitor cell

EVs:

Extracellular vesicles

IL-17:

Interleukin-17

GAS5:

Growth arrest-specific 5

SF:

Sulfasalazine

IS:

Ischemic stroke

I/R:

Ischemia–reperfusion

OGD:

Oxygen–glucose deprivation

MCAO:

Middle cerebral artery occlusion

TBI:

Traumatic brain injury

SCI:

Spinal cord injury

DHA:

Docosahexaenoic acid

MSC:

Mesenchymal stem cell

JEV:

Japanese encephalitis virus

JE:

Japanese encephalitis

HIV-1:

Human immunodeficiency virus type 1

Tat:

HIV-1 transactivator of transcription

E. coli :

Escherichia coli

S. pneumoniae :

Streptococcus pneumoniae

L. mo :

Listeria monocytogenes

M. tb :

Mycobacterium tuberculosis

PAMPs:

Pathogen-associated molecular patterns

LPS:

Lipopolysaccharide

References

  1. Engelhardt B, Vajkoczy P, Weller RO. The movers and shapers in immune privilege of the CNS. Nat Immunol. 2017;18:123–31.

    Article  CAS  PubMed  Google Scholar 

  2. Ransohoff RM, Brown MA. Innate immunity in the central nervous system. J Clin Invest. 2012;122:1164–71.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  3. Xanthos DN, Sandkuhler J. Neurogenic neuroinflammation: inflammatory CNS reactions in response to neuronal activity. Nat Rev Neurosci. 2014;15:43–53.

    Article  CAS  PubMed  Google Scholar 

  4. Yang QQ, Zhou JW. Neuroinflammation in the central nervous system: symphony of glial cells. Glia. 2019;67:1017–35.

    Article  PubMed  Google Scholar 

  5. Colonna M, Butovsky O. Microglia function in the central nervous system during health and neurodegeneration. Annu Rev Immunol. 2017;35:441–68.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  6. Orihuela R, McPherson CA, Harry GJ. Microglial M1/M2 polarization and metabolic states. Br J Pharmacol. 2016;173:649–65.

    Article  CAS  PubMed  Google Scholar 

  7. Colombo E, Farina C. Astrocytes: key regulators of neuroinflammation. Trends Immunol. 2016;37:608–20.

    Article  CAS  PubMed  Google Scholar 

  8. Sofroniew MV, Vinters HV. Astrocytes: biology and pathology. Acta Neuropathol. 2010;119:7–35.

    Article  PubMed  Google Scholar 

  9. Olabarria M, Goldman JE. Disorders of astrocytes: alexander disease as a model. Annu Rev Pathol. 2017;12:131–52.

    Article  CAS  PubMed  Google Scholar 

  10. Yang R, Huang F, Fu J, Dou B, Xu B, Miao L, Liu W, Yang X, Tan C, Chen H, Wang X. Differential transcription profiles of long non-coding RNAs in primary human brain microvascular endothelial cells in response to meningitic Escherichia coli. Sci Rep. 2016;6:38903.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  11. Consortium EP. An integrated encyclopedia of DNA elements in the human genome. Nature. 2012;489:57–74.

    Article  Google Scholar 

  12. Yang R, Xu B, Yang B, Fu J, Liu L, Amjad N, Cai A, Tan C, Chen H, Wang X. Circular RNA transcriptomic analysis of primary human brain microvascular endothelial cells infected with meningitic Escherichia coli. Mol Ther Nucleic Acids. 2018;13:651–64.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  13. Bartel DP. MicroRNAs: genomics, biogenesis, mechanism, and function. Cell. 2004;116:281–97.

    Article  CAS  PubMed  Google Scholar 

  14. Bartel DP. Metazoan microRNAs. Cell. 2018;173:20–51.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  15. Place RF, Li LC, Pookot D, Noonan EJ, Dahiya R. MicroRNA-373 induces expression of genes with complementary promoter sequences. Proc Natl Acad Sci U S A. 2008;105:1608–13.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Friedman RC, Farh KK, Burge CB, Bartel DP. Most mammalian mRNAs are conserved targets of microRNAs. Genome Res. 2009;19:92–105.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  17. Moran VA, Perera RJ, Khalil AM. Emerging functional and mechanistic paradigms of mammalian long non-coding RNAs. Nucleic Acids Res. 2012;40:6391–400.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  18. Chen LL. Linking long noncoding RNA localization and function. Trends Biochem Sci. 2016;41:761–72.

    Article  CAS  PubMed  Google Scholar 

  19. Yang R, Xu B, Yang B, Fu J, Chen H, Wang X. Non-coding RNAs: the extensive and interactive regulators of the blood-brain barrier permeability. RNA Biol. 2021;18:108–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Li X, Yang L, Chen LL. The biogenesis, functions, and challenges of circular RNAs. Mol Cell. 2018;71:428–42.

    Article  CAS  PubMed  Google Scholar 

  21. Jeck WR, Sorrentino JA, Wang K, Slevin MK, Burd CE, Liu J, Marzluff WF, Sharpless NE. Circular RNAs are abundant, conserved, and associated with ALU repeats. RNA. 2013;19:141–57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  22. Yang R, Chen J, Xu B, Yang B, Fu J, Xiao S, Tan C, Chen H, Wang X. Circ_2858 helps blood-brain barrier disruption by increasing VEGFA via sponging miR-93-5p during Escherichia coli meningitis. Mol Ther Nucleic Acids. 2020;22:708–21.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Chen LL. The expanding regulatory mechanisms and cellular functions of circular RNAs. Nat Rev Mol Cell Biol. 2020;21:475–90.

    Article  CAS  PubMed  Google Scholar 

  24. Sobue A, Komine O, Yamanaka K. Neuroinflammation in Alzheimer’s disease: microglial signature and their relevance to disease. Inflamm Regen. 2023;43:26.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Idda ML, Munk R, Abdelmohsen K, Gorospe M. Noncoding RNAs in Alzheimer’s disease. Wiley Interdiscip Rev RNA. 2018. https://doi.org/10.1002/wrna.1463.

    Article  PubMed  PubMed Central  Google Scholar 

  26. Glass CK, Saijo K, Winner B, Marchetto MC, Gage FH. Mechanisms underlying inflammation in neurodegeneration. Cell. 2010;140:918–34.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  27. Morrow A, Panyard DJ, Deming YK, Jonaitis E, Dong R, Vasiljevic E, Betthauser TJ, Kollmorgen G, Suridjan I, Bayfield A, et al. Cerebrospinal fluid sphingomyelins in Alzheimer’s disease, neurodegeneration, and neuroinflammation. J Alzheimers Dis. 2022;90:667–80.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  28. Zhai L, Shen H, Sheng Y, Guan Q. ADMSC Exo-MicroRNA-22 improve neurological function and neuroinflammation in mice with Alzheimer’s disease. J Cell Mol Med. 2021;25:7513–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  29. Nakano M, Kubota K, Kobayashi E, Chikenji TS, Saito Y, Konari N, Fujimiya M. Bone marrow-derived mesenchymal stem cells improve cognitive impairment in an Alzheimer’s disease model by increasing the expression of microRNA-146a in hippocampus. Sci Rep. 2020;10:10772.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Guedes JR, Custodia CM, Silva RJ, de Almeida LP, Pedroso de Lima MC, Cardoso AL. Early miR-155 upregulation contributes to neuroinflammation in Alzheimer’s disease triple transgenic mouse model. Hum Mol Genet. 2014;23:6286–301.

    Article  CAS  PubMed  Google Scholar 

  31. Lehmann SM, Kruger C, Park B, Derkow K, Rosenberger K, Baumgart J, Trimbuch T, Eom G, Hinz M, Kaul D, et al. An unconventional role for miRNA: let-7 activates Toll-like receptor 7 and causes neurodegeneration. Nat Neurosci. 2012;15:827–35.

    Article  CAS  PubMed  Google Scholar 

  32. Wu GD, Li ZH, Li X, Zheng T, Zhang DK. microRNA-592 blockade inhibits oxidative stress injury in Alzheimer’s disease astrocytes via the KIAA0319-mediated Keap1/Nrf2/ARE signaling pathway. Exp Neurol. 2020;324: 113128.

    Article  CAS  PubMed  Google Scholar 

  33. Chu YY, Ko CY, Wang WJ, Wang SM, Gean PW, Kuo YM, Wang JM. Astrocytic CCAAT/enhancer binding protein δ regulates neuronal viability and spatial learning ability via miR-135a. Mol Neurobiol. 2016;53:4173–88.

    Article  CAS  PubMed  Google Scholar 

  34. Xing H, Guo S, Zhang Y, Zheng Z, Wang H. Upregulation of microRNA-206 enhances lipopolysaccharide-induced inflammation and release of amyloid-beta by targeting insulin-like growth factor 1 in microglia. Mol Med Rep. 2016;14:1357–64.

    Article  CAS  PubMed  Google Scholar 

  35. Bhattacharjee S, Zhao Y, Dua P, Rogaev EI, Lukiw WJ. microRNA-34a-mediated down-regulation of the microglial-enriched triggering receptor and phagocytosis-sensor TREM2 in age-related macular degeneration. PLoS ONE. 2016;11: e0150211.

    Article  PubMed  PubMed Central  Google Scholar 

  36. Aloi MS, Prater KE, Sopher B, Davidson S, Jayadev S, Garden GA. The pro-inflammatory microRNA miR-155 influences fibrillar β-Amyloid1-42 catabolism by microglia. Glia. 2021;69:1736–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  37. Chen ML, Hong CG, Yue T, Li HM, Duan R, Hu WB, Cao J, Wang ZX, Chen CY, Hu XK, et al. Inhibition of miR-331-3p and miR-9-5p ameliorates Alzheimer’s disease by enhancing autophagy. Theranostics. 2021;11:2395–409.

    Article  PubMed  PubMed Central  Google Scholar 

  38. Boscher E, Goupil C, Petry S, Keraudren R, Loiselle A, Planel E, Hebert SS. microRNA-138 overexpression alters Aβ42 levels and behavior in wildtype mice. Front Neurosci. 2020;14: 591138.

    Article  PubMed  Google Scholar 

  39. Yi J, Chen B, Yao X, Lei Y, Ou F, Huang F. Upregulation of the lncRNA MEG3 improves cognitive impairment, alleviates neuronal damage, and inhibits activation of astrocytes in hippocampus tissues in Alzheimer’s disease through inactivating the PI3K/Akt signaling pathway. J Cell Biochem. 2019;120:18053–65.

    Article  CAS  PubMed  Google Scholar 

  40. Zheng J, Yi D, Liu Y, Wang M, Zhu Y, Shi H. Long nonding RNA UCA1 regulates neural stem cell differentiation by controlling miR-1/Hes1 expression. Am J Transl Res. 2017;9:3696–704.

    CAS  PubMed  PubMed Central  Google Scholar 

  41. Diling C, Yinrui G, Longkai Q, Xiaocui T, Yadi L, Xin Y, Guoyan H, Ou S, Tianqiao Y, Dongdong W, et al. Circular RNA NF1-419 enhances autophagy to ameliorate senile dementia by binding Dynamin-1 and Adaptor protein 2 B1 in AD-like mice. Aging (Albany NY). 2019;11:12002–31.

    Article  PubMed  Google Scholar 

  42. McGeer PL, McGeer EG. Glial reactions in Parkinson’s disease. Mov Disord. 2008;23:474–83.

    Article  PubMed  Google Scholar 

  43. Isik S, Yeman Kiyak B, Akbayir R, Seyhali R, Arpaci T. Microglia mediated neuroinflammation in Parkinson’s disease. Cells. 2023;12:1012.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  44. Pereira SL, Grossmann D, Delcambre S, Hermann A, Grunewald A. Novel insights into Parkin-mediated mitochondrial dysfunction and neuroinflammation in Parkinson’s disease. Curr Opin Neurobiol. 2023;80: 102720.

    Article  CAS  PubMed  Google Scholar 

  45. Badanjak K, Fixemer S, Smajic S, Skupin A, Grunewald A. The contribution of microglia to neuroinflammation in Parkinson’s disease. Int J Mol Sci. 2021;22:4676.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  46. Yang L, Mao K, Yu H, Chen J. Neuroinflammatory responses and Parkinson’ disease: pathogenic mechanisms and therapeutic targets. J Neuroimmune Pharmacol. 2020;15:830–7.

    Article  PubMed  Google Scholar 

  47. Li S, Bi G, Han S, Huang R. MicroRNAs play a role in Parkinson’s disease by regulating microglia function: from pathogenetic involvement to therapeutic potential. Front Mol Neurosci. 2021;14: 744942.

    Article  CAS  PubMed  Google Scholar 

  48. Thome AD, Harms AS, Volpicelli-Daley LA, Standaert DG. microRNA-155 regulates alpha-synuclein-induced inflammatory responses in models of Parkinson disease. J Neurosci. 2016;36:2383–90.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  49. He Q, Wang Q, Yuan C, Wang Y. Downregulation of miR-7116-5p in microglia by MPP+ sensitizes TNF-α production to induce dopaminergic neuron damage. Glia. 2017;65:1251–63.

    Article  PubMed  Google Scholar 

  50. Oliveira SR, Dionisio PA, Gaspar MM, Correia Guedes L, Coelho M, Rosa MM, Ferreira JJ, Amaral JD, Rodrigues CMP. miR-335 targets LRRK2 and mitigates inflammation in Parkinson’s disease. Front Cell Dev Biol. 2021;9: 661461.

    Article  PubMed  PubMed Central  Google Scholar 

  51. Yao L, Ye Y, Mao H, Lu F, He X, Lu G, Zhang S. MicroRNA-124 regulates the expression of MEKK3 in the inflammatory pathogenesis of Parkinson’s disease. J Neuroinflammation. 2018;15:13.

    Article  PubMed  PubMed Central  Google Scholar 

  52. Yao L, Zhu Z, Wu J, Zhang Y, Zhang H, Sun X, Qian C, Wang B, Xie L, Zhang S, Lu G. MicroRNA-124 regulates the expression of p62/p38 and promotes autophagy in the inflammatory pathogenesis of Parkinson’s disease. FASEB J. 2019;33:8648–65.

    Article  CAS  PubMed  Google Scholar 

  53. Lv Q, Zhong Z, Hu B, Yan S, Yan Y, Zhang J, Shi T, Jiang L, Li W, Huang W. MicroRNA-3473b regulates the expression of TREM2/ULK1 and inhibits autophagy in inflammatory pathogenesis of Parkinson disease. J Neurochem. 2021;157:599–610.

    Article  CAS  PubMed  Google Scholar 

  54. Zhou T, Lin D, Chen Y, Peng S, Jing X, Lei M, Tao E, Liang Y. α-synuclein accumulation in SH-SY5Y cell impairs autophagy in microglia by exosomes overloading miR-19a-3p. Epigenomics. 2019;11:1661–77.

    Article  CAS  PubMed  Google Scholar 

  55. Zhou Y, Lu M, Du RH, Qiao C, Jiang CY, Zhang KZ, Ding JH, Hu G. MicroRNA-7 targets Nod-like receptor protein 3 inflammasome to modulate neuroinflammation in the pathogenesis of Parkinson’s disease. Mol Neurodegener. 2016;11:28.

    Article  PubMed  PubMed Central  Google Scholar 

  56. Cao B, Wang T, Qu Q, Kang T, Yang Q. Long noncoding RNA SNHG1 promotes neuroinflammation in Parkinson’s disease via regulating miR-7/NLRP3 pathway. Neuroscience. 2018;388:118–27.

    Article  CAS  PubMed  Google Scholar 

  57. Xu W, Zhang L, Geng Y, Liu Y, Zhang N. Long noncoding RNA GAS5 promotes microglial inflammatory response in Parkinson’s disease by regulating NLRP3 pathway through sponging miR-223-3p. Int Immunopharmacol. 2020;85: 106614.

    Article  CAS  PubMed  Google Scholar 

  58. Cai LJ, Tu L, Huang XM, Huang J, Qiu N, Xie GH, Liao JX, Du W, Zhang YY, Tian JY. LncRNA MALAT1 facilitates inflammasome activation via epigenetic suppression of Nrf2 in Parkinson’s disease. Mol Brain. 2020;13:130.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  59. Cao H, Han X, Jia Y, Zhang B. Inhibition of long non-coding RNA HOXA11-AS against neuroinflammation in Parkinson’s disease model via targeting miR-124-3p mediated FSTL1/NF-κB axis. Aging (Albany NY). 2021;13:11455–69.

    Article  CAS  PubMed  Google Scholar 

  60. Ye Y, He X, Lu F, Mao H, Zhu Z, Yao L, Luo W, Sun X, Wang B, Qian C, et al. A lincRNA-p21/miR-181 family feedback loop regulates microglial activation during systemic LPS- and MPTP- induced neuroinflammation. Cell Death Dis. 2018;9:803.

    Article  PubMed  PubMed Central  Google Scholar 

  61. Akcimen F, Lopez ER, Landers JE, Nath A, Chio A, Chia R, Traynor BJ. Amyotrophic lateral sclerosis: translating genetic discoveries into therapies. Nat Rev Genet 2023.

  62. Merjane J, Chung R, Patani R, Lisowski L. Molecular mechanisms of amyotrophic lateral sclerosis as broad therapeutic targets for gene therapy applications utilizing adeno-associated viral vectors. Med Res Rev. 2023;43:829–54.

    Article  CAS  PubMed  Google Scholar 

  63. Peggion C, Scalcon V, Massimino ML, Nies K, Lopreiato R, Rigobello MP, Bertoli A. SOD1 in ALS: taking stock in pathogenic mechanisms and the role of glial and muscle cells. Antioxidants (Basel). 2022;11:614.

    Article  CAS  PubMed  Google Scholar 

  64. Cunha C, Santos C, Gomes C, Fernandes A, Correia AM, Sebastiao AM, Vaz AR, Brites D. Downregulated glia interplay and increased miRNA-155 as promising markers to track ALS at an early stage. Mol Neurobiol. 2018;55:4207–24.

    CAS  PubMed  Google Scholar 

  65. Butovsky O, Jedrychowski MP, Cialic R, Krasemann S, Murugaiyan G, Fanek Z, Greco DJ, Wu PM, Doykan CE, Kiner O, et al. Targeting miR-155 restores abnormal microglia and attenuates disease in SOD1 mice. Ann Neurol. 2015;77:75–99.

    Article  CAS  PubMed  Google Scholar 

  66. Parisi C, Napoli G, Amadio S, Spalloni A, Apolloni S, Longone P, Volonte C. MicroRNA-125b regulates microglia activation and motor neuron death in ALS. Cell Death Differ. 2016;23:531–41.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  67. Barbosa M, Gomes C, Sequeira C, Goncalves-Ribeiro J, Pina CC, Carvalho LA, Moreira R, Vaz SH, Vaz AR, Brites D. Recovery of depleted miR-146a in ALS cortical astrocytes reverts cell aberrancies and prevents paracrine pathogenicity on microglia and motor neurons. Front Cell Dev Biol. 2021;9: 634355.

    Article  PubMed  PubMed Central  Google Scholar 

  68. Zhou F, Zhang C, Guan Y, Chen Y, Lu Q, Jie L, Gao H, Du H, Zhang H, Liu Y, Wang X. Screening the expression characteristics of several miRNAs in G93A-SOD1 transgenic mouse: altered expression of miRNA-124 is associated with astrocyte differentiation by targeting Sox2 and Sox9. J Neurochem. 2018;145:51–67.

    Article  CAS  PubMed  Google Scholar 

  69. Ng W, Ng SY. Remodeling of astrocyte secretome in amyotrophic lateral sclerosis: uncovering novel targets to combat astrocyte-mediated toxicity. Transl Neurodegener. 2022;11:54.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  70. Varcianna A, Myszczynska MA, Castelli LM, O’Neill B, Kim Y, Talbot J, Nyberg S, Nyamali I, Heath PR, Stopford MJ, et al. Micro-RNAs secreted through astrocyte-derived extracellular vesicles cause neuronal network degeneration in C9orf72 ALS. EBioMedicine. 2019;40:626–35.

    Article  PubMed  PubMed Central  Google Scholar 

  71. Hoye ML, Regan MR, Jensen LA, Lake AM, Reddy LV, Vidensky S, Richard JP, Maragakis NJ, Rothstein JD, Dougherty JD, Miller TM. Motor neuron-derived microRNAs cause astrocyte dysfunction in amyotrophic lateral sclerosis. Brain. 2018;141:2561–75.

    PubMed  PubMed Central  Google Scholar 

  72. Koutsouraki E, Theodoros K, Eleni G, Marianna K, Areti N, Ariadni K, Dimitrios M. Autonomic nervous system disorders in multiple sclerosis. J Neurol. 2023;270:3703.

    Article  PubMed  Google Scholar 

  73. Baecher-Allan C, Kaskow BJ, Weiner HL. Multiple sclerosis: mechanisms and immunotherapy. Neuron. 2018;97:742–68.

    Article  CAS  PubMed  Google Scholar 

  74. das Neves SP, Sousa JC, Sousa N, Cerqueira JJ, Marques F. Altered astrocytic function in experimental neuroinflammation and multiple sclerosis. Glia. 2021;69:1341–68.

    Article  PubMed  Google Scholar 

  75. Junker A, Krumbholz M, Eisele S, Mohan H, Augstein F, Bittner R, Lassmann H, Wekerle H, Hohlfeld R, Meinl E. MicroRNA profiling of multiple sclerosis lesions identifies modulators of the regulatory protein CD47. Brain. 2009;132:3342–52.

    Article  PubMed  Google Scholar 

  76. Moore CS, Rao VT, Durafourt BA, Bedell BJ, Ludwin SK, Bar-Or A, Antel JP. miR-155 as a multiple sclerosis-relevant regulator of myeloid cell polarization. Ann Neurol. 2013;74:709–20.

    Article  CAS  PubMed  Google Scholar 

  77. Tarassishin L, Loudig O, Bauman A, Shafit-Zagardo B, Suh HS, Lee SC. Interferon regulatory factor 3 inhibits astrocyte inflammatory gene expression through suppression of the proinflammatory miR-155 and miR-155*. Glia. 2011;59:1911–22.

    Article  PubMed  PubMed Central  Google Scholar 

  78. Lopez-Ramirez MA, Wu D, Pryce G, Simpson JE, Reijerkerk A, King-Robson J, Kay O, de Vries HE, Hirst MC, Sharrack B, et al. MicroRNA-155 negatively affects blood-brain barrier function during neuroinflammation. FASEB J. 2014;28:2551–65.

    Article  CAS  PubMed  Google Scholar 

  79. Zhang J, Zhang ZG, Lu M, Zhang Y, Shang X, Chopp M. MiR-146a promotes oligodendrocyte progenitor cell differentiation and enhances remyelination in a model of experimental autoimmune encephalomyelitis. Neurobiol Dis. 2019;125:154–62.

    Article  CAS  PubMed  Google Scholar 

  80. Prada I, Gabrielli M, Turola E, Iorio A, D’Arrigo G, Parolisi R, De Luca M, Pacifici M, Bastoni M, Lombardi M, et al. Glia-to-neuron transfer of miRNAs via extracellular vesicles: a new mechanism underlying inflammation-induced synaptic alterations. Acta Neuropathol. 2018;135:529–50.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  81. Li Y, Zhou D, Ren Y, Zhang Z, Guo X, Ma M, Xue Z, Lv J, Liu H, Xi Q, et al. Mir223 restrains autophagy and promotes CNS inflammation by targeting ATG16L1. Autophagy. 2019;15:478–92.

    Article  CAS  PubMed  Google Scholar 

  82. Galloway DA, Blandford SN, Berry T, Williams JB, Stefanelli M, Ploughman M, Moore CS. miR-223 promotes regenerative myeloid cell phenotype and function in the demyelinated central nervous system. Glia. 2019;67:857–69.

    Article  PubMed  Google Scholar 

  83. Ponomarev ED, Veremeyko T, Barteneva N, Krichevsky AM, Weiner HL. MicroRNA-124 promotes microglia quiescence and suppresses EAE by deactivating macrophages via the C/EBP-α-PU.1 pathway. Nat Med. 2011;17:64–70.

    Article  CAS  PubMed  Google Scholar 

  84. Fang X, Sun D, Wang Z, Yu Z, Liu W, Pu Y, Wang D, Huang A, Liu M, Xiang Z, et al. miR-30a positively regulates the inflammatory response of microglia in experimental autoimmune encephalomyelitis. Neurosci Bull. 2017;33:603–15.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  85. Khan D, Ansar Ahmed S. Regulation of IL-17 in autoimmune diseases by transcriptional factors and microRNAs. Front Genet. 2015;6:236.

    Article  PubMed  PubMed Central  Google Scholar 

  86. Zhu S, Pan W, Song X, Liu Y, Shao X, Tang Y, Liang D, He D, Wang H, Liu W, et al. The microRNA miR-23b suppresses IL-17-associated autoimmune inflammation by targeting TAB2, TAB3 and IKK-α. Nat Med. 2012;18:1077–86.

    Article  CAS  PubMed  Google Scholar 

  87. Liu X, He F, Pang R, Zhao D, Qiu W, Shan K, Zhang J, Lu Y, Li Y, Wang Y. Interleukin-17 (IL-17)-induced microRNA 873 (miR-873) contributes to the pathogenesis of experimental autoimmune encephalomyelitis by targeting A20 ubiquitin-editing enzyme. J Biol Chem. 2014;289:28971–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  88. Liu X, Zhou F, Yang Y, Wang W, Niu L, Zuo D, Li X, Hua H, Zhang B, Kou Y, et al. miR-409-3p and miR-1896 co-operatively participate in IL-17-induced inflammatory cytokine production in astrocytes and pathogenesis of EAE mice via targeting SOCS3/STAT3 signaling. Glia. 2019;67:101–12.

    Article  PubMed  Google Scholar 

  89. Liu X, Zhou F, Wang W, Chen G, Zhang Q, Lv R, Zhao Z, Li X, Yu Q, Meves JM, et al. IL-9-triggered lncRNA Gm13568 regulates Notch1 in astrocytes through interaction with CBP/P300: contribute to the pathogenesis of experimental autoimmune encephalomyelitis. J Neuroinflammation. 2021;18:108.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  90. Sun D, Yu Z, Fang X, Liu M, Pu Y, Shao Q, Wang D, Zhao X, Huang A, Xiang Z, et al. LncRNA GAS5 inhibits microglial M2 polarization and exacerbates demyelination. EMBO Rep. 2017;18:1801–16.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  91. Duan C, Liu Y, Li Y, Chen H, Liu X, Chen X, Yue J, Zhou X, Yang J. Sulfasalazine alters microglia phenotype by competing endogenous RNA effect of miR-136-5p and long non-coding RNA HOTAIR in cuprizone-induced demyelination. Biochem Pharmacol. 2018;155:110–23.

    Article  CAS  PubMed  Google Scholar 

  92. Waseem A, Rashid S, Rashid K, Khan MA, Khan R, Haque R, Seth P, Raza SS. Insight into the transcription factors regulating ischemic stroke and glioma in response to shared stimuli. Semin Cancer Biol. 2023;92:102–27.

    Article  CAS  PubMed  Google Scholar 

  93. Zhang Z, Zou X, Zhang R, Xie Y, Feng Z, Li F, Han J, Sun H, Ouyang Q, Hua S, et al. Human umbilical cord mesenchymal stem cell-derived exosomal miR-146a-5p reduces microglial-mediated neuroinflammation via suppression of the IRAK1/TRAF6 signaling pathway after ischemic stroke. Aging (Albany NY). 2021;13:3060–79.

    Article  CAS  PubMed  Google Scholar 

  94. Xu S, Lu J, Shao A, Zhang JH, Zhang J. Glial cells: role of the immune response in ischemic stroke. Front Immunol. 2020;11:294.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  95. Candelario-Jalil E, Dijkhuizen RM, Magnus T. Neuroinflammation, stroke, blood-brain barrier dysfunction, and imaging modalities. Stroke. 2022;53:1473–86.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  96. Kadir RRA, Alwjwaj M, Bayraktutan U. MicroRNA: an emerging predictive, diagnostic, prognostic and therapeutic strategy in ischaemic stroke. Cell Mol Neurobiol. 2022;42:1301–19.

    Article  PubMed  Google Scholar 

  97. Lian L, Zhang Y, Liu L, Yang L, Cai Y, Zhang J, Xu S. Neuroinflammation in ischemic stroke: focus on microRNA-mediated polarization of microglia. Front Mol Neurosci. 2020;13: 612439.

    Article  CAS  PubMed  Google Scholar 

  98. Kong H, Omran A, Ashhab MU, Gan N, Peng J, He F, Wu L, Deng X, Yin F. Changes in microglial inflammation-related and brain-enriched microRNAs expressions in response to in vitro oxygen-glucose deprivation. Neurochem Res. 2014;39:233–43.

    Article  CAS  PubMed  Google Scholar 

  99. Xu X, Wen Z, Zhao N, Xu X, Wang F, Gao J, Jiang Y, Liu X. MicroRNA-1906, a novel regulator of toll-like receptor 4, ameliorates ischemic injury after experimental stroke in mice. J Neurosci. 2017;37:10498–515.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  100. Kolosowska N, Gotkiewicz M, Dhungana H, Giudice L, Giugno R, Box D, Huuskonen MT, Korhonen P, Scoyni F, Kanninen KM, et al. Intracerebral overexpression of miR-669c is protective in mouse ischemic stroke model by targeting MyD88 and inducing alternative microglial/macrophage activation. J Neuroinflammation. 2020;17:194.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  101. Hamzei Taj S, Kho W, Riou A, Wiedermann D, Hoehn M. MiRNA-124 induces neuroprotection and functional improvement after focal cerebral ischemia. Biomaterials. 2016;91:151–65.

    Article  CAS  PubMed  Google Scholar 

  102. Song Y, Li Z, He T, Qu M, Jiang L, Li W, Shi X, Pan J, Zhang L, Wang Y, et al. M2 microglia-derived exosomes protect the mouse brain from ischemia-reperfusion injury via exosomal miR-124. Theranostics. 2019;9:2910–23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  103. Li Z, Song Y, He T, Wen R, Li Y, Chen T, Huang S, Wang Y, Tang Y, Shen F, et al. M2 microglial small extracellular vesicles reduce glial scar formation via the miR-124/STAT3 pathway after ischemic stroke in mice. Theranostics. 2021;11:1232–48.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  104. Zhao H, Wang J, Gao L, Wang R, Liu X, Gao Z, Tao Z, Xu C, Song J, Ji X, Luo Y. MiRNA-424 protects against permanent focal cerebral ischemia injury in mice involving suppressing microglia activation. Stroke. 2013;44:1706–13.

    Article  CAS  PubMed  Google Scholar 

  105. Zhao H, Li G, Wang R, Tao Z, Zhang S, Li F, Han Z, Li L, Liu P, Luo Y. MiR-424 prevents astrogliosis after cerebral ischemia/reperfusion in elderly mice by enhancing repressive H3K27me3 via NFIA/DNMT1 signaling. FEBS J. 2019;286:4926–36.

    Article  CAS  PubMed  Google Scholar 

  106. Ni J, Wang X, Chen S, Liu H, Wang Y, Xu X, Cheng J, Jia J, Zhen X. MicroRNA let-7c-5p protects against cerebral ischemia injury via mechanisms involving the inhibition of microglia activation. Brain Behav Immun. 2015;49:75–85.

    Article  CAS  PubMed  Google Scholar 

  107. Yang J, Cao LL, Wang XP, Guo W, Guo RB, Sun YQ, Xue TF, Cai ZY, Ji J, Cheng H, Sun XL. Neuronal extracellular vesicle derived miR-98 prevents salvageable neurons from microglial phagocytosis in acute ischemic stroke. Cell Death Dis. 2021;12:23.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  108. Bernstein DL, Zuluaga-Ramirez V, Gajghate S, Reichenbach NL, Polyak B, Persidsky Y, Rom S. miR-98 reduces endothelial dysfunction by protecting blood-brain barrier (BBB) and improves neurological outcomes in mouse ischemia/reperfusion stroke model. J Cereb Blood Flow Metab. 2020;40:1953–65.

    Article  CAS  PubMed  Google Scholar 

  109. Ma F, Sun P, Zhang X, Hamblin MH, Yin KJ. Endothelium-targeted deletion of the miR-15a/16–1 cluster ameliorates blood-brain barrier dysfunction in ischemic stroke. Sci Signal. 2020;13:eaay5686.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  110. Wang X, Chen S, Ni J, Cheng J, Jia J, Zhen X. miRNA-3473b contributes to neuroinflammation following cerebral ischemia. Cell Death Dis. 2018;9:11.

    Article  PubMed  PubMed Central  Google Scholar 

  111. Ma Q, Zhao H, Tao Z, Wang R, Liu P, Han Z, Ma S, Luo Y, Jia J. MicroRNA-181c exacerbates brain injury in acute ischemic stroke. Aging Dis. 2016;7:705–14.

    Article  PubMed  PubMed Central  Google Scholar 

  112. Wang J, Chen T, Shan G. miR-148b regulates proliferation and differentiation of neural stem cells via Wnt/β-catenin signaling in rat ischemic stroke model. Front Cell Neurosci. 2017;11:329.

    Article  PubMed  PubMed Central  Google Scholar 

  113. Mo JL, Liu Q, Kou ZW, Wu KW, Yang P, Chen XH, Sun FY. MicroRNA-365 modulates astrocyte conversion into neuron in adult rat brain after stroke by targeting Pax6. Glia. 2018;66:1346–62.

    Article  PubMed  PubMed Central  Google Scholar 

  114. Liu Z, Zhou Y, Xia J. CircRNAs: key molecules in the prevention and treatment of ischemic stroke. Biomed Pharmacother. 2022;156: 113845.

    Article  CAS  PubMed  Google Scholar 

  115. Vasudeva K, Dutta A, Munshi A. Role of lncRNAs in the development of ischemic stroke and their therapeutic potential. Mol Neurobiol. 2021;58:3712–28.

    Article  CAS  PubMed  Google Scholar 

  116. Zhang X, Zhu XL, Ji BY, Cao X, Yu LJ, Zhang Y, Bao XY, Xu Y, Jin JL. LncRNA-1810034E14Rik reduces microglia activation in experimental ischemic stroke. J Neuroinflammation. 2019;16:75.

    Article  PubMed  PubMed Central  Google Scholar 

  117. Ni X, Su Q, Xia W, Zhang Y, Jia K, Su Z, Li G. Knockdown lncRNA NEAT1 regulates the activation of microglia and reduces AKT signaling and neuronal apoptosis after cerebral ischemic reperfusion. Sci Rep. 2020;10:19658.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  118. Wang J, Zhao H, Fan Z, Li G, Ma Q, Tao Z, Wang R, Feng J, Luo Y. Long noncoding RNA H19 promotes neuroinflammation in ischemic stroke by driving histone deacetylase 1-dependent M1 microglial polarization. Stroke. 2017;48:2211–21.

    Article  CAS  PubMed  Google Scholar 

  119. Wang H, Liao S, Li H, Chen Y, Yu J. Long non-coding RNA TUG1 sponges miR-145a-5p to regulate microglial polarization after oxygen-glucose deprivation. Front Mol Neurosci. 2019;12:215.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  120. Zhang H, Lu M, Zhang X, Kuai Y, Mei Y, Tan Q, Zhong K, Sun X, Tan W. Isosteviol sodium protects against ischemic stroke by modulating microglia/macrophage polarization via disruption of GAS5/miR-146a-5p sponge. Sci Rep. 2019;9:12221.

    Article  PubMed  PubMed Central  Google Scholar 

  121. Wang H, Zheng X, Jin J, Zheng L, Guan T, Huo Y, Xie S, Wu Y, Chen W. LncRNA MALAT1 silencing protects against cerebral ischemia-reperfusion injury through miR-145 to regulate AQP4. J Biomed Sci. 2020;27:40.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  122. Deng Y, Chen D, Wang L, Gao F, Jin B, Lv H, Zhang G, Sun X, Liu L, Mo D, et al. Silencing of long noncoding RNA Nespas aggravates microglial cell death and neuroinflammation in ischemic stroke. Stroke. 2019;50:1850–8.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  123. Chen J, Jin J, Zhang X, Yu H, Zhu X, Yu L, Chen Y, Liu P, Dong X, Cao X, et al. Microglial lnc-U90926 facilitates neutrophil infiltration in ischemic stroke via MDH2/CXCL2 axis. Mol Ther. 2021;29:2873–85.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  124. 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;21:806–15.

    Article  CAS  PubMed  Google Scholar 

  125. Han B, Zhang Y, Zhang Y, Bai Y, Chen X, Huang R, Wu F, Leng S, Chao J, Zhang JH, et al. Novel insight into circular RNA HECTD1 in astrocyte activation via autophagy by targeting MIR142-TIPARP: implications for cerebral ischemic stroke. Autophagy. 2018;14:1164–84.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  126. Chen W, Wang H, Zhu Z, Feng J, Chen L. Exosome-shuttled circSHOC2 from IPASs regulates neuronal autophagy and ameliorates ischemic brain injury via the miR-7670-3p/SIRT1 axis. Mol Ther Nucleic Acids. 2020;22:657–72.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  127. Dams-O’Connor K, Juengst SB, Bogner J, Chiaravalloti ND, Corrigan JD, Giacino JT, Harrison-Felix CL, Hoffman JM, Ketchum JM, Lequerica AH, et al. Traumatic brain injury as a chronic disease: insights from the United States traumatic brain injury model systems research program. Lancet Neurol. 2023;22:517–28.

    Article  PubMed  Google Scholar 

  128. Karve IP, Taylor JM, Crack PJ. The contribution of astrocytes and microglia to traumatic brain injury. Br J Pharmacol. 2016;173:692–702.

    Article  CAS  PubMed  Google Scholar 

  129. Korotkov A, Puhakka N, Gupta SD, Vuokila N, Broekaart DWM, Anink JJ, Heiskanen M, Karttunen J, van Scheppingen J, Huitinga I, et al. Increased expression of miR142 and miR155 in glial and immune cells after traumatic brain injury may contribute to neuroinflammation via astrocyte activation. Brain Pathol. 2020;30:897–912.

    CAS  PubMed  PubMed Central  Google Scholar 

  130. Kumar A, Stoica BA, Loane DJ, Yang M, Abulwerdi G, Khan N, Kumar A, Thom SR, Faden AI. Microglial-derived microparticles mediate neuroinflammation after traumatic brain injury. J Neuroinflammation. 2017;14:47.

    Article  PubMed  PubMed Central  Google Scholar 

  131. Henry RJ, Doran SJ, Barrett JP, Meadows VE, Sabirzhanov B, Stoica BA, Loane DJ, Faden AI. Inhibition of miR-155 limits neuroinflammation and improves functional recovery after experimental traumatic brain injury in mice. Neurotherapeutics. 2019;16:216–30.

    Article  CAS  PubMed  Google Scholar 

  132. Jadhav SP, Kamath SP, Choolani M, Lu J, Dheen ST. microRNA-200b modulates microglia-mediated neuroinflammation via the cJun/MAPK pathway. J Neurochem. 2014;130:388–401.

    Article  CAS  PubMed  Google Scholar 

  133. Mao S, Li X, Wang J, Ding X, Zhang C, Li L. miR-17-92 facilitates neuronal differentiation of transplanted neural stem/precursor cells under neuroinflammatory conditions. J Neuroinflammation. 2016;13:208.

    Article  PubMed  PubMed Central  Google Scholar 

  134. Wu J, Li H, He J, Tian X, Luo S, Li J, Li W, Zhong J, Zhang H, Huang Z, et al. Downregulation of microRNA-9-5p promotes synaptic remodeling in the chronic phase after traumatic brain injury. Cell Death Dis. 2021;12:9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  135. Mohamadzadeh O, Hajinouri M, Moammer F, Tamehri Zadeh SS, Omid Shafiei G, Jafari A, Ostadian A, Talaei Zavareh SA, Hamblin MR, Yazdi AJ, et al. Non-coding RNAs and exosomal non-coding RNAs in traumatic brain injury: the small player with big actions. Mol Neurobiol. 2023;60:4064–83.

    Article  CAS  PubMed  Google Scholar 

  136. Ge X, Guo M, Hu T, Li W, Huang S, Yin Z, Li Y, Chen F, Zhu L, Kang C, et al. Increased microglial exosomal miR-124-3p alleviates neurodegeneration and improves cognitive outcome after rmTBI. Mol Ther. 2020;28:503–22.

    Article  CAS  PubMed  Google Scholar 

  137. Huang S, Ge X, Yu J, Han Z, Yin Z, Li Y, Chen F, Wang H, Zhang J, Lei P. Increased miR-124-3p in microglial exosomes following traumatic brain injury inhibits neuronal inflammation and contributes to neurite outgrowth via their transfer into neurons. FASEB J. 2018;32:512–28.

    Article  CAS  PubMed  Google Scholar 

  138. Yin Z, Han Z, Hu T, Zhang S, Ge X, Huang S, Wang L, Yu J, Li W, Wang Y, et al. Neuron-derived exosomes with high miR-21-5p expression promoted polarization of M1 microglia in culture. Brain Behav Immun. 2020;83:270–82.

    Article  CAS  PubMed  Google Scholar 

  139. Long X, Yao X, Jiang Q, Yang Y, He X, Tian W, Zhao K, Zhang H. Astrocyte-derived exosomes enriched with miR-873a-5p inhibit neuroinflammation via microglia phenotype modulation after traumatic brain injury. J Neuroinflammation. 2020;17:89.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  140. Zhang L, Wang H. Long Non-coding RNA in CNS injuries: a new target for therapeutic intervention. Mol Ther Nucleic Acids. 2019;17:754–66.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  141. Meng J, Ding T, Chen Y, Long T, Xu Q, Lian W, Liu W. LncRNA-Meg3 promotes Nlrp3-mediated microglial inflammation by targeting miR-7a-5p. Int Immunopharmacol. 2021;90: 107141.

    Article  CAS  PubMed  Google Scholar 

  142. Liu N, Sun H, Li X, Cao W, Peng A, Dong S, Yu Z. Downregulation of lncRNA KCNQ1OT1 relieves traumatic brain injury induced neurological deficits via promoting “M2” microglia polarization. Brain Res Bull. 2021;171:91–102.

    Article  CAS  PubMed  Google Scholar 

  143. Cheng S, Zhang Y, Chen S, Zhou Y. LncRNA HOTAIR participates in microglia activation and inflammatory factor release by regulating the ubiquitination of MYD88 in traumatic brain injury. J Mol Neurosci. 2021;71:169–77.

    Article  CAS  PubMed  Google Scholar 

  144. Yu Y, Cao F, Ran Q, Wang F. Long non-coding RNA Gm4419 promotes trauma-induced astrocyte apoptosis by targeting tumor necrosis factor α. Biochem Biophys Res Commun. 2017;491:478–85.

    Article  CAS  PubMed  Google Scholar 

  145. Zhang Y, Wang J, Zhang Y, Wei J, Wu R, Cai H. Overexpression of long noncoding RNA Malat1 ameliorates traumatic brain injury induced brain edema by inhibiting AQP4 and the NF-κB/IL-6 pathway. J Cell Biochem. 2019;120:17584–92.

    Article  CAS  PubMed  Google Scholar 

  146. He B, Chen W, Zeng J, Tong W, Zheng P. Long noncoding RNA NKILA transferred by astrocyte-derived extracellular vesicles protects against neuronal injury by upregulating NLRX1 through binding to mir-195 in traumatic brain injury. Aging (Albany NY). 2021;13:8127–45.

    Article  CAS  PubMed  Google Scholar 

  147. Kim HN, McCrea MR, Li S. Advances in molecular therapies for targeting pathophysiology in spinal cord injury. Expert Opin Ther Targets. 2023;27:171–87.

    Article  CAS  PubMed  Google Scholar 

  148. Ahuja CS, Wilson JR, Nori S, Kotter MRN, Druschel C, Curt A, Fehlings MG. Traumatic spinal cord injury. Nat Rev Dis Primers. 2017;3:17018.

    Article  PubMed  Google Scholar 

  149. Li X, Li M, Tian L, Chen J, Liu R, Ning B. Reactive astrogliosis: implications in spinal cord injury progression and therapy. Oxid Med Cell Longev. 2020;2020:9494352.

    Article  PubMed  PubMed Central  Google Scholar 

  150. Hong P, Jiang M, Li H. Functional requirement of dicer1 and miR-17-5p in reactive astrocyte proliferation after spinal cord injury in the mouse. Glia. 2014;62:2044–60.

    Article  PubMed  Google Scholar 

  151. Wang CY, Yang SH, Tzeng SF. MicroRNA-145 as one negative regulator of astrogliosis. Glia. 2015;63:194–205.

    Article  PubMed  Google Scholar 

  152. Liu R, Wang W, Wang S, Xie W, Li H, Ning B. microRNA-21 regulates astrocytic reaction post-acute phase of spinal cord injury through modulating TGF-β signaling. Aging (Albany NY). 2018;10:1474–88.

    Article  CAS  PubMed  Google Scholar 

  153. Silvestro S, Mazzon E. MiRNAs as promising translational strategies for neuronal repair and regeneration in spinal cord injury. Cells. 2022;11:2177.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  154. Louw AM, Kolar MK, Novikova LN, Kingham PJ, Wiberg M, Kjems J, Novikov LN. Chitosan polyplex mediated delivery of miRNA-124 reduces activation of microglial cells in vitro and in rat models of spinal cord injury. Nanomedicine. 2016;12:643–53.

    Article  CAS  PubMed  Google Scholar 

  155. Jiang D, Gong F, Ge X, Lv C, Huang C, Feng S, Zhou Z, Rong Y, Wang J, Ji C, et al. Neuron-derived exosomes-transmitted miR-124-3p protect traumatically injured spinal cord by suppressing the activation of neurotoxic microglia and astrocytes. J Nanobiotechnology. 2020;18:105.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  156. Yip PK, Bowes AL, Hall JCE, Burguillos MA, Ip THR, Baskerville T, Liu ZH, Mohamed M, Getachew F, Lindsay AD, et al. Docosahexaenoic acid reduces microglia phagocytic activity via miR-124 and induces neuroprotection in rodent models of spinal cord contusion injury. Hum Mol Genet. 2019;28:2427–48.

    Article  CAS  PubMed  Google Scholar 

  157. Liu W, Rong Y, Wang J, Zhou Z, Ge X, Ji C, Jiang D, Gong F, Li L, Chen J, et al. Exosome-shuttled miR-216a-5p from hypoxic preconditioned mesenchymal stem cells repair traumatic spinal cord injury by shifting microglial M1/M2 polarization. J Neuroinflammation. 2020;17:47.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  158. Qian Z, Chang J, Jiang F, Ge D, Yang L, Li Y, Chen H, Cao X. Excess administration of miR-340-5p ameliorates spinal cord injury-induced neuroinflammation and apoptosis by modulating the P38-MAPK signaling pathway. Brain Behav Immun. 2020;87:531–42.

    Article  CAS  PubMed  Google Scholar 

  159. Vivinetto AL, Kim ID, Goldberg DC, Fones L, Brown E, Tarabykin VS, Hill CE, Cho S, Cave JW. Zeb2 is a regulator of astrogliosis and functional recovery after CNS injury. Cell Rep. 2020;31: 107834.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  160. Wei H, Wu X, You Y, Duran RC, Zheng Y, Narayanan KL, Hai B, Li X, Tallapragada N, Prajapati TJ, et al. Systematic analysis of purified astrocytes after SCI unveils Zeb2os function during astrogliosis. Cell Rep. 2021;34: 108721.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  161. Shao M, Jin M, Xu S, Zheng C, Zhu W, Ma X, Lv F. Exosomes from long noncoding RNA-Gm37494-ADSCs repair spinal cord injury via shifting microglial M1/M2 polarization. Inflammation. 2020;43:1536–47.

    Article  CAS  PubMed  Google Scholar 

  162. Zhou HJ, Wang LQ, Wang DB, Yu JB, Zhu Y, Xu QS, Zheng XJ, Zhan RY. Long noncoding RNA MALAT1 contributes to inflammatory response of microglia following spinal cord injury via the modulation of a miR-199b/IKKβ/NF-κB signaling pathway. Am J Physiol Cell Physiol. 2018;315:C52–61.

    Article  CAS  PubMed  Google Scholar 

  163. Zhao Q, Lu F, Su Q, Liu Z, Xia X, Yan Z, Zhou F, Qin R. Knockdown of long noncoding RNA XIST mitigates the apoptosis and inflammatory injury of microglia cells after spinal cord injury through miR-27a/Smurf1 axis. Neurosci Lett. 2020;715: 134649.

    Article  CAS  PubMed  Google Scholar 

  164. Xiang W, Jiang L, Zhou Y, Li Z, Zhao Q, Wu T, Cao Y, Zhou J. The lncRNA Ftx/miR-382-5p/Nrg1 axis improves the inflammation response of microglia and spinal cord injury repair. Neurochem Int. 2021;143: 104929.

    Article  CAS  PubMed  Google Scholar 

  165. Cui SY, Zhang W, Cui ZM, Yi H, Xu DW, Liu W, Zhu XH. Knockdown of long non-coding RNA LEF1-AS1 attenuates apoptosis and inflammatory injury of microglia cells following spinal cord injury. J Orthop Surg Res. 2021;16:6.

    Article  PubMed  PubMed Central  Google Scholar 

  166. Xia X, Niu H, Ma Y, Qu B, He M, Yu K, Wang E, Zhang L, Gu J, Liu G. LncRNA CCAT1 protects astrocytes against OGD/R-induced damage by targeting the miR-218/NFAT5-signaling axis. Cell Mol Neurobiol. 2020;40:1383–93.

    Article  CAS  PubMed  Google Scholar 

  167. Xu S, Wang J, Jiang J, Song J, Zhu W, Zhang F, Shao M, Xu H, Ma X, Lyu F. TLR4 promotes microglial pyroptosis via lncRNA-F630028O10Rik by activating PI3K/AKT pathway after spinal cord injury. Cell Death Dis. 2020;11:693.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  168. Jiang ZS, Zhang JR. LncRNA SNHG5 enhances astrocytes and microglia viability via upregulating KLF4 in spinal cord injury. Int J Biol Macromol. 2018;120:66–72.

    Article  CAS  PubMed  Google Scholar 

  169. Wu R, Mao S, Wang Y, Zhou S, Liu Y, Liu M, Gu X, Yu B. Differential circular RNA expression profiles following spinal cord injury in rats: a temporal and experimental analysis. Front Neurosci. 2019;13:1303.

    Article  PubMed  PubMed Central  Google Scholar 

  170. Tohidpour A, Morgun AV, Boitsova EB, Malinovskaya NA, Martynova GP, Khilazheva ED, Kopylevich NV, Gertsog GE, Salmina AB. Neuroinflammation and infection: molecular mechanisms associated with dysfunction of neurovascular unit. Front Cell Infect Microbiol. 2017;7:276.

    Article  PubMed  PubMed Central  Google Scholar 

  171. Yang RC, Huang K, Zhang HP, Li L, Zhang YF, Tan C, Chen HC, Jin ML, Wang XR. SARS-CoV-2 productively infects human brain microvascular endothelial cells. J Neuroinflammation. 2022;19:149.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  172. Yang RC, Huang K, Zhang HP, Li L, Tan C, Chen HC, Jin ML, Wang XR. Transcriptional landscape of human neuroblastoma cells in response to SARS-CoV-2. BMC Neurosci. 2022;23:43.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  173. Zhu B, Ye J, Nie Y, Ashraf U, Zohaib A, Duan X, Fu ZF, Song Y, Chen H, Cao S. MicroRNA-15b modulates Japanese encephalitis virus-mediated inflammation via targeting RNF125. J Immunol. 2015;195:2251–62.

    Article  CAS  PubMed  Google Scholar 

  174. Deng M, Du G, Zhao J, Du X. miR-146a negatively regulates the induction of proinflammatory cytokines in response to Japanese encephalitis virus infection in microglial cells. Arch Virol. 2017;162:1495–505.

    Article  CAS  PubMed  Google Scholar 

  175. Sharma N, Verma R, Kumawat KL, Basu A, Singh SK. miR-146a suppresses cellular immune response during Japanese encephalitis virus JaOArS982 strain infection in human microglial cells. J Neuroinflammation. 2015;12:30.

    Article  PubMed  PubMed Central  Google Scholar 

  176. Thounaojam MC, Kundu K, Kaushik DK, Swaroop S, Mahadevan A, Shankar SK, Basu A. MicroRNA 155 regulates Japanese encephalitis virus-induced inflammatory response by targeting Src homology 2-containing inositol phosphatase 1. J Virol. 2014;88:4798–810.

    Article  PubMed  PubMed Central  Google Scholar 

  177. Pareek S, Roy S, Kumari B, Jain P, Banerjee A, Vrati S. MiR-155 induction in microglial cells suppresses Japanese encephalitis virus replication and negatively modulates innate immune responses. J Neuroinflammation. 2014;11:97.

    Article  PubMed  PubMed Central  Google Scholar 

  178. Thounaojam MC, Kaushik DK, Kundu K, Basu A. MicroRNA-29b modulates Japanese encephalitis virus-induced microglia activation by targeting tumor necrosis factor alpha-induced protein 3. J Neurochem. 2014;129:143–54.

    Article  CAS  PubMed  Google Scholar 

  179. Hazra B, Chakraborty S, Bhaskar M, Mukherjee S, Mahadevan A, Basu A. miR-301a regulates inflammatory response to Japanese encephalitis virus infection via suppression of NKRF activity. J Immunol. 2019;203:2222–38.

    Article  CAS  PubMed  Google Scholar 

  180. Mukherjee S, Akbar I, Kumari B, Vrati S, Basu A, Banerjee A. Japanese encephalitis virus-induced let-7a/b interacted with the NOTCH-TLR7 pathway in microglia and facilitated neuronal death via caspase activation. J Neurochem. 2019;149:518–34.

    Article  CAS  PubMed  Google Scholar 

  181. Ashraf U, Zhu B, Ye J, Wan S, Nie Y, Chen Z, Cui M, Wang C, Duan X, Zhang H, et al. MicroRNA-19b-3p modulates Japanese encephalitis virus-mediated inflammation via targeting RNF11. J Virol. 2016;90:4780–95.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  182. Jadhav VS, Krause KH, Singh SK. HIV-1 Tat C modulates NOX2 and NOX4 expressions through miR-17 in a human microglial cell line. J Neurochem. 2014;131:803–15.

    Article  CAS  PubMed  Google Scholar 

  183. Thompson KA, Cherry CL, Bell JE, McLean CA. Brain cell reservoirs of latent virus in presymptomatic HIV-infected individuals. Am J Pathol. 2011;179:1623–9.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  184. Periyasamy P, Thangaraj A, Bendi VS, Buch S. HIV-1 Tat-mediated microglial inflammation involves a novel miRNA-34a-NLRC5-NFκB signaling axis. Brain Behav Immun. 2019;80:227–37.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  185. Periyasamy P, Thangaraj A, Guo ML, Hu G, Callen S, Buch S. Epigenetic promoter DNA methylation of miR-124 promotes HIV-1 Tat-mediated microglial activation via MECP2-STAT3 axis. J Neurosci. 2018;38:5367–83.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  186. Rom S, Rom I, Passiatore G, Pacifici M, Radhakrishnan S, Del Valle L, Pina-Oviedo S, Khalili K, Eletto D, Peruzzi F. CCL8/MCP-2 is a target for mir-146a in HIV-1-infected human microglial cells. FASEB J. 2010;24:2292–300.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  187. Thangaraj A, Chivero ET, Tripathi A, Singh S, Niu F, Guo ML, Pillai P, Periyasamy P, Buch S. HIV TAT-mediated microglial senescence: role of SIRT3-dependent mitochondrial oxidative stress. Redox Biol. 2021;40: 101843.

    Article  CAS  PubMed  Google Scholar 

  188. Fatima M, Prajapati B, Saleem K, Kumari R, Mohindar Singh Singal C, Seth P. Novel insights into role of miR-320a-VDAC1 axis in astrocyte-mediated neuronal damage in neuroAIDS. Glia. 2017;65:250–63.

    Article  PubMed  Google Scholar 

  189. Rahimian P, He JJ. HIV-1 Tat-shortened neurite outgrowth through regulation of microRNA-132 and its target gene expression. J Neuroinflammation. 2016;13:247.

    Article  PubMed  PubMed Central  Google Scholar 

  190. Hu G, Yao H, Chaudhuri AD, Duan M, Yelamanchili SV, Wen H, Cheney PD, Fox HS, Buch S. Exosome-mediated shuttling of microRNA-29 regulates HIV Tat and morphine-mediated neuronal dysfunction. Cell Death Dis. 2012;3: e381.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  191. Pilakka-Kanthikeel S, Raymond A, Atluri VS, Sagar V, Saxena SK, Diaz P, Chevelon S, Concepcion M, Nair M. Sterile alpha motif and histidine/aspartic acid domain-containing protein 1 (SAMHD1)-facilitated HIV restriction in astrocytes is regulated by miRNA-181a. J Neuroinflammation. 2015;12:66.

    Article  PubMed  PubMed Central  Google Scholar 

  192. Chao TC, Zhang Q, Li Z, Tiwari SK, Qin Y, Yau E, Sanchez A, Singh G, Chang K, Kaul M, et al. The long noncoding RNA HEAL regulates HIV-1 replication through epigenetic regulation of the HIV-1 promoter. MBio. 2019;10:e02016-e2019.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  193. Sil S, Hu G, Liao K, Niu F, Callen S, Periyasamy P, Fox HS, Buch S. HIV-1 Tat-mediated astrocytic amyloidosis involves the HIF-1α/lncRNA BACE1-AS axis. PLoS Biol. 2020;18: e3000660.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  194. Yang R, Wang J, Wang F, Zhang H, Tan C, Chen H, Wang X. Blood-brain barrier integrity damage in bacterial meningitis: the underlying link, mechanisms, and therapeutic targets. Int J Mol Sci. 2023;24:2852.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  195. Geyer S, Jacobs M, Hsu NJ. Immunity against bacterial infection of the central nervous system: an astrocyte perspective. Front Mol Neurosci. 2019;12:57.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  196. Yang RC, Qu XY, Xiao SY, Li L, Xu BJ, Fu JY, Lv YJ, Amjad N, Tan C, Kim KS, et al. Meningitic Escherichia coli-induced upregulation of PDGF-B and ICAM-1 aggravates blood-brain barrier disruption and neuroinflammatory response. J Neuroinflammation. 2019;16:101.

    Article  PubMed  PubMed Central  Google Scholar 

  197. Yang R, Lv Y, Miao L, Zhang H, Qu X, Chen J, Xu B, Yang B, Fu J, Tan C, et al. Resveratrol attenuates meningitic Escherichia coli-mediated blood-brain barrier disruption. ACS Infect Dis. 2021;7:777–89.

    Article  CAS  PubMed  Google Scholar 

  198. Barichello T, Generoso JS, Simoes LR, Goularte JA, Petronilho F, Saigal P, Badawy M, Quevedo J. Role of microglial activation in the pathophysiology of bacterial meningitis. Mol Neurobiol. 2016;53:1770–81.

    Article  CAS  PubMed  Google Scholar 

  199. Karthikeyan A, Patnala R, Jadhav SP, Eng-Ang L, Dheen ST. MicroRNAs: key players in microglia and astrocyte mediated inflammation in CNS pathologies. Curr Med Chem. 2016;23:3528–46.

    Article  CAS  PubMed  Google Scholar 

  200. Yang B, Yin P, Yang R, Xu B, Fu J, Zhi S, Dai M, Tan C, Chen H, Wang X. Holistic insights into meningitic Escherichia coli infection of astrocytes based on whole transcriptome profiling. Epigenomics. 2020;12:1611–32.

    Article  CAS  PubMed  Google Scholar 

  201. Yang B, Yang R, Xu B, Fu J, Qu X, Li L, Dai M, Tan C, Chen H, Wang X. miR-155 and miR-146a collectively regulate meningitic Escherichia coli infection-mediated neuroinflammatory responses. J Neuroinflammation. 2021;18:114.

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  202. Feng X, Hu J, Zhan F, Luo D, Hua F, Xu G. MicroRNA-138-5p regulates hippocampal neuroinflammation and cognitive impairment by NLRP3/caspase-1 signaling pathway in rats. J Inflamm Res. 2021;14:1125–43.

    Article  PubMed  PubMed Central  Google Scholar 

  203. Feng X, Zhan F, Luo D, Hu J, Wei G, Hua F, Xu G. LncRNA 4344 promotes NLRP3-related neuroinflammation and cognitive impairment by targeting miR-138-5p. Brain Behav Immun. 2021;98:283–98.

    Article  CAS  PubMed  Google Scholar 

  204. Zhang M, Gillaspy AF, Gipson JR, Cassidy BR, Nave JL, Brewer MF, Stoner JA, Chen J, Drevets DA. Neuroinvasive Listeria monocytogenes infection triggers IFN-activation of microglia and upregulates microglial miR-155. Front Immunol. 2018;9:2751.

    Article  PubMed  PubMed Central  Google Scholar 

  205. Jin Y, Yao G, Wang Y, Teng L, Wang Y, Chen H, Gao R, Lin W, Wang Z, Chen J. MiR-30c-5p mediates inflammatory responses and promotes microglia survival by targeting eIF2α during Cryptococcus neoformans infection. Microb Pathog. 2020;141: 103959.

    Article  CAS  PubMed  Google Scholar 

  206. Zhou YL, Zhang L, Zhou Z, Liu W, Lu Y, He S, Cui Y, Qin Y, Hua M. Antibody modified nanoparticle-mediated delivery of miR-124 regulates apoptosis via repression the Stat3 signal in Mycobacterial-infected microglia. J Biomed Nanotechnol. 2018;14:2185–97.

    Article  CAS  PubMed  Google Scholar 

  207. Fang X, Wang H, Zhuo Z, Tian P, Chen Z, Wang Y, Cheng X. miR-141-3p inhibits the activation of astrocytes and the release of inflammatory cytokines in bacterial meningitis through down-regulating HMGB1. Brain Res. 2021;1770: 147611.

    Article  CAS  PubMed  Google Scholar 

  208. Dong Y, Wang J, Du KX, Jia TM, Zhu CL, Zhang Y, Xu FL. MicroRNA-135a participates in the development of astrocytes derived from bacterial meningitis by downregulating HIF-1α. Am J Physiol Cell Physiol. 2019;316:C711–21.

    Article  CAS  PubMed  Google Scholar 

  209. Zhang TM, Ma SH, Lv JK, Wang XY, Afewerky HK, Li H, Lu YM. The emerging role of exosomes in Alzheimer’s disease. Ageing Res Rev. 2021;68: 101321.

    Article  CAS  PubMed  Google Scholar 

Download references

Acknowledgements

Not applicable.

Funding

This work was supported by the National Key Research and Development Program of China (2021YFD1800800), the National Natural Science Foundation of China (NSFC) (32122086 and 32102749), the China Postdoctoral Science Foundation (2022M721277), Natural Science Foundation of Hubei Province (2021CFA016), and the Fundamental Research Funds for the Central Universities (2662023PY005).

Author information

Authors and Affiliations

Authors

Contributions

Conceptualization and written, RY, BY, and XW; supervision, XW, CT, and HC; review and editing, XW, RY, WL; funding acquisition, XW and RY. Both authors have read and agreed to the published version of the manuscript.

Corresponding author

Correspondence to Xiangru Wang.

Ethics declarations

Ethics approval and consent to participate

Not applicable.

Consent for publication

Not applicable.

Competing interests

The authors declare that they have no competing interests.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Rights and permissions

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/. The Creative Commons Public Domain Dedication waiver (http://creativecommons.org/publicdomain/zero/1.0/) applies to the data made available in this article, unless otherwise stated in a credit line to the data.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Yang, R., Yang, B., Liu, W. et al. Emerging role of non-coding RNAs in neuroinflammation mediated by microglia and astrocytes. J Neuroinflammation 20, 173 (2023). https://doi.org/10.1186/s12974-023-02856-0

Download citation

  • Received:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1186/s12974-023-02856-0

Keywords