IMR Press / JIN / Volume 23 / Issue 4 / DOI: 10.31083/j.jin2304087
Open Access Review
The Roles of Circular RNAs in Ischemic Stroke through Modulating Neuroinflammation
Show Less
1 The Fourth School of Clinical Medicine, Zhejiang Chinese Medical University, 310053 Hangzhou, Zhejiang, China
2 Department of Neurology, Hangzhou First People’s Hospital, 310006 Hangzhou, Zhejiang, China
3 Department of Cardiology, Hangzhou First People’s Hospital, 310006 Hangzhou, Zhejiang, China
4 Department of Neurosurgery, the Second Affiliated Hospital, School of Medicine, Zhejiang University, 310009 Hangzhou, Zhejiang, China
5 Key Laboratory of Precise Treatment and Clinical Translational Research of Neurological Disease, 310012 Hangzhou, Zhejiang, China
*Correspondence: jianglin0229@126.com (Lin Jiang); shaoanwen@zju.edu.cn (Anwen Shao); yincg716@aliyun.com (Congguo Yin)
These authors contributed equally.
J. Integr. Neurosci. 2024, 23(4), 87; https://doi.org/10.31083/j.jin2304087
Submitted: 13 November 2023 | Revised: 3 December 2023 | Accepted: 12 December 2023 | Published: 23 April 2024
Copyright: © 2024 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Ischemic stroke (IS) remains a serious threat to human health. Neuroinflammatory response is an important pathophysiological process after IS. Circular RNAs (circRNAs), a member of the non-coding RNA family, are highly expressed in the central nervous system and widely involved in regulating physiological and pathophysiological processes. This study reviews the current evidence on neuroinflammatory responses, the role of circRNAs in IS and their potential mechanisms in regulating inflammatory cells, and inflammatory factors affecting IS damage. This review lays a foundation for future clinical application of circRNAs as novel biomarkers and therapeutic targets.

Keywords
ischemic stroke
neuroinflammation
circRNA
inflammatory cytokine
therapeutic target
1. Introduction

According to the Global Burden of Disease Study 2019, stroke is the second leading cause of level 3 death worldwide in 2019, after ischemic heart disease, and the third leading cause of level 3 death and disability, posing a serious threat to human health and social development [1]. Ischemic stroke (IS) is a clinical syndrome characterized by cerebral tissue ischemia, hypoxia, and necrosis resulting from arterial narrowing or occlusion in the brain’s blood supply (carotid and vertebral arteries) [1, 2]. In 2019, IS accounted for 62.4% of all new strokes worldwide, making it the predominant type of stroke [1].

Following IS, brain inflammation is the major pathological event causing secondary cerebral tissue injury and poor functional recovery [3, 4, 5, 6, 7]. Post-stroke inflammation is initiated after the initial ischemia, marked by innate immune responses and the breakdown of the blood-brain barrier (BBB), releasing proinflammatory cytokines and infiltrating immune cells. Additionally, many studies on neuroinflammation have revealed that glial cells regulate the release of inflammatory cytokines and the progression of the inflammatory cascade [8, 9].

Circular RNAs (circRNAs), belonging to the non-coding RNA family, have a stable structure and high tissue-specific expression. CircRNAs are highly expressed and widely involved in regulating physiological and pathophysiological processes in the nervous system [10, 11].

CircRNAs are involved in neuronal injury and repair, glial cell protection, and BBB injury after IS [12, 13, 14, 15, 16, 17, 18]. However, the potential roles of different circRNAs in inflammatory responses remain unclear. Therefore, this review aims to provide new insights into the role of circRNAs in modulating neuroinflammatory responses in IS to enhance our understanding of its severity and prognosis.

2. Pathological Mechanism of Neuroinflammation after IS
2.1 Inflammatory Role of Immune Cells

Currently, excitotoxicity, calcium dysregulation, oxidative and nitrosative stress, cortical spreading depolarizations, and inflammation are used to explain the pathophysiology of IS [19, 20]. Ischemic neuroinflammation is a pathological process of IS that significantly affects clinical prognosis [21]. Notably, cytokines, secreted by different brain immune cells, are key players in the inflammatory response, exhibiting both proinflammatory and anti-inflammatory effects [22, 23]. Fig. 1 shows the inflammatory response of immune cells.

Fig. 1.

Pathophysiological process of neuroinflammatory response after ischemic stroke. After ischemic brain injury, DAMPs are released during cell ischemia, hypoxia, and necrosis, promoting the activation of microglia, astrocytes, brain endothelial cells, and perivascular immune cells. Microglia polarize into classical (M1) or alternative (M2) microglia and secrete proinflammatory cytokines, anti-inflammatory cytokines, adhesion molecules, and chemokines. Activated astrocytes initially secrete proinflammatory cytokines, followed by anti-inflammatory cytokines, chemokines, and adhesion molecules. Proinflammatory cytokines contribute to the BBB breakdown and exacerbate the inflammatory response, while anti-inflammatory cytokines counteract this process and play a neuroprotective role. Adhesion molecules and chemokines facilitate the infiltration of peripheral immune cells into the BBB and promote the inflammatory response. DAMPs, damage-associated molecular patterns; BBB, blood-brain barrier; TLRs, Toll-like receptors; M1, classical microglia; M2, alternative microglia.

After ischemic brain injury, ischemia-anoxic necrosis cells release a series of danger signals commonly called damage-associated molecular patterns (DAMPs), such as the high-mobility group box 1 protein (HMGB1) [24, 25, 26]. These DAMPs are released into the extracellular environment to activate the innate immune system and further stimulate the inflammatory cascade reaction [27]. Toll-like receptors (TLRs), located on microglia, astrocytes, perivascular macrophages, and brain endothelial cells, detect DAMPs and activate corresponding cells in response [28, 29, 30, 31].

Microglia, the brain’s most predominant resident macrophages, account for approximately 10% of all brain tissue cells that become highly activated early after stroke [32, 33, 34]. After activation, microglia polarize into two types, classical (M1) and alternative (M2) microglia, which play different roles, secreting cytokines that promote or relieve inflammation and affecting tissue repair [35, 36]. This process increases microglial phagocytosis activity [37]. Currently, microglial cells secrete a variety of inflammatory cytokines and chemokines, including interleukin(IL)-1β, IL-4, IL-6, IL-8, IL-10, IL-12, IL-15, tumor necrosis factor (TNF)-α, transforming growth factor (TGF)-β, macrophage inflammatory protein (MIP)-1, monocyte chemoattractant protein-1 (MCP-1), chemokines interferon-γ-inducible protein 10 (CXCL10, also called IP-10), and proteolytic enzymes including matrix metalloproteinase (MMP)-3 and MMP-9 [38, 39, 40].

In the central nervous system, astrocytes are the main type of glial cells, accounting for approximately 50% of brain cells. In addition to playing a supporting role, astrocytes are important regulatory factors of brain function [41, 42, 43]. When activated by proinflammatory factors secreted by activated microglia, astrocytes lose their neuroprotective and homeostatic ability and induce neuron and oligodendrocyte death [44]. After brain tissue ischemia, astrocytes reduce brain nerve damage by producing antioxidants and glutamine and regulating extracellular potassium concentration [45]. Activated astrocytes secrete large amounts of inflammatory cytokines and chemokines, such as IL-1, TNF-α, interferon-γ (IFN-γ), and MCP-1, which contribute to BBB destruction [46, 47]. While astrocyte death leads to decreased expression of tight junction proteins, leading to BBB breakdown and brain tissue edema [48]. Inflammatory factors such as IL-1, TNF-α, and MCP-1 increase in the early stages of cerebral ischemic injury, while IL-6 and macrophage migration inhibitory factors (MIF) increase at later stages [49]. Moreover, proinflammatory factors enhance the expression of adhesion molecules, including E-selectin, P-selectin, intercellular cell adhesion molecule (ICAM)-1, ICAM-2, and vascular endothelial cell adhesion molecule (VCAM)-1 [50, 51, 52, 53].

Neutrophils are one of the peripheral immune cells. Neutrophils also release cytokines that promote glial scarring, phagocytic debris, and edema resolution [54]. Following BBB damage, neutrophils penetrate the brain through the action of cell adhesion molecules, including selectin, integrin, and immunoglobulin [55]. Neutrophils have a strong ability to destroy brain tissue, as they generate reactive oxygen species (ROS) and proteolytic enzymes, inducing direct neurotoxic effects. Additionally, their accumulation indirectly leads to cerebral ischemia and hypoxia by blocking capillary blood flow [56, 57].

Another peripheral cells, (such as mononuclear phagocytes, natural killer cells (NK cells), and lymphocytes) also release cytokines and participate in ischemic inflammation [23, 58]. Monocytes are innate immune cells that serve as important effectors and regulatory factors in inflammation [59], crossing the BBB, accumulating in damaged brain tissue, and exhibiting microglia-like phenotypic changes [60, 61]. NK cells are important innate immune cells in peripheral blood that mobilize quickly in the initial stage of inflammatory responses, thereby promoting neuroinflammatory and inducing neuronal death [62].

A growing body of literature has begun to elucidate that lymphocytes can assume in IS pathology. Lymphocytes cross the BBB and enter ischemic brain parenchyma under the influence of P-selectin, VCAM-1, and ICAM-1 signaling [63]. The degree of lymphocyte infiltration depends on changes in BBB permeability [64]. Immune system activation mediated by lymphocyte infiltration can have both harmful antigen-specific autoreactive responses and cellular protection in IS [65]. During this process, helper T cells are divided into Th1 cells secreting proinflammatory cytokines (such as IFN-γ, TNF-α, and lymphotoxin alpha (LT-α)), Th2 cells secreting anti-inflammatory cytokines (such as IL-4 and IL-10), and TH17 cells secreting IL-17, IL-21, and IL-22 cytokines [66]. γδ T cells, associated with post-stroke neurotoxicity, are induced by macrophage-secreted IL-23 and DAMPs-activated surface TLRs [67]. Activated γδ T cells migrate to the ischemic tissue and produce IFN-γ and IL-17 [68]. IFN-γ, mainly secreted by T cells, mediates delayed neurotoxic effects in IS [69, 70]. Regulatory T cells (Treg cells) exhibit stronger anti-apoptotic properties compared to other cells [71] and secrete cytokines such as IL-10, which have protective effects after cerebral ischemia. Treg cells reduce the early invasion and activation of neutrophils and T cells. When Treg cells are depleted, the expressions of TNF-α, IL-1β, IFN-γ, and other proinflammatory cytokines increase in the brain [72]. Regulatory B cells also have beneficial effects on the ischemic brain by secreting IL-10 [73]. Studies have shown that lack of B cells increases migration cross the BBB and decreases IL-10 production. Conversely, infiltration of B cells into the brain reduces brain infarct size and peripheral T cell activation [73].

2.2 Inflammatory Action of Cytokines and Chemicals

Brain ischemia triggers a surge in inflammatory cytokine production by various cell types, including microglia, astrocytes, endothelial cells, and neurons, with contrasting outcomes. Proinflammatory cytokines (such as TNF-α and IL-1) exacerbate stroke-related brain injury, while anti-inflammatory cytokines (such as IL-10 and TGF-β) alleviate cerebral ischemia injury [23, 55, 74]. Table 1 (Ref. [8, 22, 55, 69, 75, 76, 77, 78, 79, 80, 81, 82, 83, 84, 85, 86, 87, 88, 89, 90, 91, 92, 93, 94, 95, 96, 97, 98, 99, 100, 101, 102, 103, 104, 105, 106, 107, 108, 109, 110, 111, 112, 113, 114, 115, 116, 117, 118, 119, 120, 121, 122, 123, 124]) outlines the role of inflammatory mediators.

Table 1.The inflammatory action of cytokines and chemicals after ischemic stroke.
Classification Designation Produced by Roles in inflammatory response References
Pro-inflammatory cytokines TNF-α Neurons, microglia, astrocytes, monocytes, Th1 cells, and endothelial cells Combine TNF receptors, form complex1, active caspase, and induce apoptosis [22, 75, 76, 77]
Upregulate nuclear factor kappa-B (NF-κB) activation, increase glutamate, NO, MMPs, and ROS
Activate microglia and astrocytes and increase infarct volume
Increase chemokines and adhesion molecules (ICAM and VCAM), promote coagulation and BBB breakdown, promote neutrophil infiltration
Enhance the toxic effects of IL-1
IL-1α/β Neurons, microglia, astrocytes, monocytes, and endothelial cells Increase NMDAR function and Ca2+ concentration and lead to neurogenic death [55, 78, 79, 80, 81]
Promote the expression of adhesion molecules (ICAM-1, ICAM-2, P-selectin, E-selectin, and VCAM-1) and induce BBB breakdown
Activate lipoprotein-associated phospholipase 2 (LP-A2), induce apoptosis, release metabolites (such as prostaglandins and leukotrienes), induce BBB breakdown, and promote leukocyte infiltration
Activate microglia and astrocytes and increase IL-1, IL-6, TNF-α, NO, iNOS, and MMP levels
Upregulate NF-κB activation
IL-12 Microglia, monocytes, NK cells, and T cells Induce inflammatory T cell and increase infarct volume [8, 82]
IL-15 Neurons, microglia, macrophages, astrocytes, and endothelial cells Promote the migration of T cells [83, 84, 85]
Activate NK cells and increase infarct volumes
IL-17 Neurons, microglia, astrocytes, monocytes, endothelial cells, Th17 cells, γδ T cells, and NK cells Upregulate NF-κB activation, induce BBB breakdown, and promote neutrophil infiltration [86]
IL-21 Th17 cells, B cells, and NK cells Activate the janus kinase/signal transducer and activator of transcription (JAK/STAT) pathway [87, 88]
Increase infarct size and lymphocyte accumulation
IL-23 Monocytes, microglia, and astrocytes Promote T-cell activation to produce IL-17 [89]
IFN-γ Astrocyte, (γδ) T cells, NK cells, microglia, and macrophages Increase the expression of IP-10 and infarct size [69, 90]
Anti-inflammatory cytokines IL-4 Th2 cells, microglia, and monocytes Polarize microglia, promote angiogenesis, alleviate ischemic stroke [91, 92]
Downregulate the expression of TNF-α and IL-1β
IL-6 Neurons, microglia, astrocytes, monocytes, endothelial cells, lymphocytes, T cells, B cells, and neutrophils Promote tissue recovery, reduce infarct severity, induce neutrophil apoptosis, inhibit the expression of TNF-α, and induce the synthesis of IL-1 receptor antagonists [93, 94, 95, 96]
IL-10 Neurons, microglia, astrocytes, monocytes, endothelial cells, Th cells, Treg cells, and B cells Reduce infarct size and inhibit cell apoptosis [97, 98]
Increase nerve growth factor (NGF) and l-glutathione (GSH) and decrease IL-1β, TNF-α, IFN-γ levels
IL-22 Th17 cells Activate the JAK2/STAT3 pathway [99]
Reduce infarct size and inhibit cell apoptosis
Decrease IL-1β, TNF-α, and MCP-1 levels
TGF-β Microglia, astrocytes, and macrophages Promote glial scar formation and microglia polarizing [100, 101]
Inhibit apoptosis, counteract excitotoxic neuronal loss, and reduce infarct size and neuroinflammation
IGF-1 Neurons, astrocytes, endothelial cells, and macrophages Promote microglia polarizing and decrease IL-1β, TNF-α, iNOS, and ROS levels [102]
Maintain astrocyte glutamate uptake and glucose homeostasis
Reduce neurogenic death
Chemokines IL-8 Microglia, monocytes, endothelial cells, and T cells Induce neutrophil adhesion and migration [103, 104]
MIF Astrocyte Promote leukocyte recruitment and thrombus formation [105]
MIP-1α/β Microglia Promote macrophage and neutrophil migration, increase inflammatory mediators, and increase infarct size [106]
MCP-1 Neurons, microglia, macrophages, astrocytes, and endothelial cells Induce BBB breakdown, promote macrophage migration [106]
Increase infarct size
IP-10 Neurons, microglia, astrocytes, monocytes, endothelial cells, and NK cells Induce NK cell accumulation and BBB breakdown [107]
Free oxygen radicals ROS Neurons, microglia, neutrophil, and astrocytes Activate the NF-ĸβ pathway, cause inflammatory cascades, damage mitochondria, and induce autophagy [108, 109]
NO Neurons, macrophages, astrocytes, microglia, and endothelial cells Induce mitochondrial dysfunction and energy depletion, promote angiogenesis, and increase infarct size [110, 111, 112]
Produce neurotoxic effect and promote inflammation, cell death, and BBB damage
iNOS Endothelial cells, astrocytes, microglia, and neutrophils Promote NO production and vasodilation, and increase infarct volume [113, 114]
Protease MMPs (MMP-3/9) Neurons, microglia, astrocytes, neutrophil, endothelial cells, and macrophages Induce BBB breakdown and neuronal apoptosis [115, 116]
Promote leukocyte adherence and transmigration
Induce cerebral edema and hemorrhagic complications
Increase infarct size
Adhesion molecules ICAMs (ICAM-1, ICAM-4) Microglia, astrocytes, macrophages, endothelial cells, and leukocytes Promote neutrophils’ entry into the BBB [117, 118]
VCAM-1 Microglia, astrocytes, macrophages, endothelial cells, and leukocytes Promote macrophages’ entry into the BBB and neovascularization [119, 120]
P-selectin Microglia, astrocytes, macrophages, endothelial cells, and leukocytes Promote leukocyte adhesion and migration [121]
E-selectin Microglia, astrocytes, macrophages, endothelial cells, and leukocytes Induce upregulation of endothelial cell adhesion molecules [122, 123, 124]
Promote leukocyte adhesion and migration

TNF-α, tumor necrosis factor-α; IL, interleukin; IFN-γ, interferon-γ; TGF-β, transforming growth factor-β; IGF-1, insulin-like growth factor 1; MIF, migration inhibitory factor; MIP-1α/β, macrophage inflammatory protein-1α/β; MCP-1, monocyte chemoattractant protein-1; IP-10, Interferon-γ (IFN-γ)-induced protein 10; ROS, reactive oxygen species; NO, nitric oxide; iNOS, inducible nitric oxide synthase; MMPs, matrix metalloproteinases; ICAMs, intercellular cell adhesion molecules; VCAM-1, vascular endothelial cell adhesion molecule-1; NF-κB, nuclear factor kappa-B; NMDAR, N-Methyl-D-aspartic acid receptor; BBB, blood-brain barrier; JAK/STAT, janus kinase/signal transducer and activator of transcription; NK cells, natural killer cells; NGF, nerve growth factor; GSH, glutathione.

Previous studies have associated differences in the production of inflammatory to infarct size following cerebral ischemia. Proinflammatory factors can increase cell edema and death, while the expression of anti-inflammatory factors can partially alleviate nerve damage [125, 126]. IL-1 is widely produced in nerve cells, microglia, and astrocytes after brain tissue ischemia. It is an important proinflammatory cytokine in the inflammatory response [55]. IL-1 can promote neuronal death by inducing calcium ions to enter cells by acting on N-Methyl-D-aspartic acid receptors (NMDARs) [78]. IL-1 also activates phospholipase A2, disrupting cell phospholipase bilayers, inducing apoptosis, and promoting the release of metabolites such as prostaglandins and leukotrienes, which can induce BBB breakdown and promote inflammatory response [93, 127]. Furthermore, IL-1 induces BBB breakdown and promotes leukocyte infiltration through adhesion molecules on endothelial cells [55, 128, 129, 130]. Adhesion molecules act as ligands involved in leukocyte chemotaxis and inflammatory responses by binding to leukocyte integrin receptors attached to endothelial cells [20]. Moreover, inflammatory stimulation results in a significant upregulation of ICAM-1 and animal studies have shown that artificial intervention with IL-1 receptor antagonists can reduce cerebral infarct size [131, 132, 133].

IL-6 is a proinflammatory cytokine that is highly expressed after a stroke event [134] and is associated with brain tissue necrosis and poor clinical prognosis [55, 94, 135]. However, there is another view that IL-6 has an anti-inflammatory effect by promoting tissue recovery and reducing infarct severity [136, 137, 138]. Currently, the known anti-inflammatory action mechanisms include inhibiting inflammatory factors, inhibiting TNF-α expression, inducing the production of IL-1 receptor antagonists, and inducing neutrophil apoptosis [95, 139]. Therefore, the role of IL-6 in the neuroinflammatory pathophysiological of IS is complex.

TNF-α is an important proinflammatory cytokine produced by microglia, monocytes, and other immune cells during acute inflammation, resulting in cell necrosis or apoptosis through intracellular signaling pathways [140]. TNF-α level was positively correlated with cerebral infarction volume [141, 142]. Artificial use of anti-TNF-α and TNF-α binding protein could decrease tissue necrosis and infarct size [143, 144, 145, 146]. TNF-α can activate microglia and astrocytes, induce the expression of adhesion molecules in brain endothelial cells, destroy the BBB, promote leukocyte infiltration, and affect glutamate transfer and synaptic plasticity [22, 75].

IL-10 is an anti-inflammatory cytokine widely present in astrocytes, mononuclear macrophages, Th cells, and B cells, and animal experiments have demonstrated that mice with IL-10 overexpression have smaller infarct size and limited apoptosis [97]. Previous research showed that IL-10 can reduce infarct size after focal ischemia, reduce leukocyte and monocyte/macrophage infiltration, and mitigate nerve damage caused by inflammatory cytokines [72, 147, 148].

TGF-β, a neuroprotective factor, promotes tissue regeneration and neurological recovery [55, 149, 150]. Experimental results showed that TGF-β can promote endothelial cell angiogenesis, stimulate M2 polarization of microglia cells, and inhibit neuron damage [151]. Moreover, TGF-β stabilizes the BBB during early stroke [152, 153].

Nuclear factor kappa-B (NF-κB), a heterogeneous transcription factor activated after cerebral ischemia, induces inflammation [154, 155, 156]. Studies have shown that NF-κB influences microglia activation and polarization [157]while promoting the expression of inflammatory cytokines, MMPs, and adhesion molecules in inflammatory response [158, 159, 160]. Multiple pathways, including the suppressor of cytokine signaling 1/janus kinase/signal transducer and activator of transcription (SOCS-1/Janus/STAT) signaling pathway, are involved [161].

MMPs are a family of proteolytic enzymes which can induce BBB breakdown [162]. During the acute phase of IS, MMP expression increases [163, 164]. MMPs are found in microglia, astrocytes, and endothelial cells. Moreover, MMPs activate microglia, macrophages, and neutrophils, further promoting MMP secretion and leading to vascular and BBB damage [165]. Experimental studies have shown that silencing MMP-Ⅸ and MMP-Ⅲ gene expression significantly alleviated BBB damage and edema [166, 167].

3. Features of circRNA
Discovery, Functions, Biogenesis and Classification of circRNAs

CircRNAs are non-coding RNA molecules that form covalent ring structures without 5 terminal cap or 3 terminal poly (A) tails [168]. CircRNAs were first discovered in plant viroids in 1976 and were widespread in the cytoplasm of eukaryotes [169, 170, 171, 172, 173, 174, 175]. Initially considered as by-products of mis-splicing [176], circRNAs currently have various functions (Fig. 2), including (i) acting as microRNA (miRNA) sponges: circRNAs contain miRNA binding sites, competitively binding miRNAs to indirectly regulate their transcription and translation [177]; (ii) impacting RNA polymerase II transcription: circRNAs can form R-loops with their producing locus or coactivate transcription factors (TFs) to modulate transcription [178, 179]; (iii) regulating splicing: circRNA back-splicing sites can compete with pre-mRNA splicing, influencing mRNA synthesis [180, 181]; (iv) absorbing or combining proteins: circRNAs containing protein-binding sites can bind to proteins, influencing their participation in other neuronal functions [180, 182, 183, 184]; (v) forming functional circRNA-protein (circRNP) complexes: several circRNAs can form circRNP complexes with RNA binding proteins (RBPs), thereby modulating signaling pathways [185, 186, 187]; (vi) affecting mRNA expression: partial circRNAs can impact mRNAs’ stability and translation by directly binding to mRNAs [188, 189]; and (vii) protein translation: some circRNAs through splicing-dependent and cap-independent mechanisms, can be translated into proteins [190, 191, 192].

Fig. 2.

Biogenesis and function of circRNAs. (A) Lariat-driven circularization model: the 3 splice donor of exon 1 and the 5 splice acceptor of exon 4 links up end-to-end by exon skipping and form an exon-containing lariat structure. Finally, the ecircRNA forms after intron removal. (B) Intron pairing-driven circularization model: direct base pairing of introns forms a circulation structure, thereby forming ecircRNA or EIciRNA after intron removal. (C) RBP-dependent cyclization model: RBPs bridge two flanking introns close together and then remove introns to form circRNAs. (D) ciRNA formation model: the elements near the splice site escape debranching stably to form the intron lariat from the splicing reaction. (E) tRNA intronic circRNA: tricRNAs derived from introns removed during precursor-tRNA (pre-tRNA) splicing. (F) Act as microRNA (miRNA) sponges. (G) Impact RNA Pol II transcription. (H) Regulate splicing. (I) Absorb or combine proteins. (J) Form functional circRNP complexes. (K) Interact with mRNAs to affect their expression. circRNAs, Circular RNAs; ecircRNAs, exonic circRNAs; EIciRNAs, exon-intron circRNAs; RBP, RNA binding protein; ciRNAs, circular intronic RNAs; tricRNAs, tRNA intronic circular RNAs.

CircRNAs can be categorized into four groups based on their biogenesis: exonic circRNAs (ecircRNAs), synthesized from only the exon sequence, circular intronic RNAs (ciRNAs) containing introns, exon-intron circRNAs (EIciRNAs) containing both exon and intron sequences, and tRNA intronic circular RNAs (tricRNAs) [10, 173, 193, 194, 195]. The biogenesis of circRNAs is illustrated in Fig. 2.

4. Role of circRNAs in IS
4.1 Initiative circRNAs were Involved in the Inflammatory Process of IS
4.1.1 CircRNAs were Abnormally Expressed in Brain Tissue and Peripheral Blood in Different Animal Models

The earliest study on circRNA expression profiles was published in STROKE in 2017 by Mehta et al. [196]. Using a transient middle cerebral artery occlusion (tMCAO) mice model, they detected 1320 circRNAs from exonic gene regions in the ischemic penumbral cortex. Among these, 283 were significantly altered after tMCAO compared to controls. This abnormal circRNA expression may play a potential pathophysiological role in the ischemic brain tissues of mice with tMCAO. In another study, Duan et al. [197] identified 14,694 differentially expressed circRNAs in MCAO rat brain tissues through high-throughput sequencing and quantitative reverse transcription-polymerase chain reaction (RT-PCR), with 87 differentially expressed circRNAs displaying significant fold changes. Moreover, Liu et al. [198] used RNA extraction and microarray assays to analyze circRNA expression in reperfused mice brain tissues after transient ischemia, revealing differential regulation of 1027 circRNAs, with 914 upregulated and 113 downregulated. These studies suggest the presence of abnormal circRNA expression in ischemic brain tissue, with potential pathophysiological implications after MCAO.

In addition to the ischemic core and penumbra, significantly expressed circRNAs were also identified in non-ischemic areas and peripheral blood from an MCAO mouse model. In one experiment, focal cortical ischemia was induced in adult male distal MCAO mice, and circRNA expression in the ipsilateral non-ischemic thalamic regions was measured using high throughput sequencing. The results revealed 2659 circRNAs with significant alterations in the non-infarct area of the ipsilateral hemisphere [199]. Another experiment validated the differential expression of blood circRNAs in tMCAO mice, confirming their role after IS. The result showed 10739 highly expressed circRNAs, and their expression across different time points showed no significant difference, indicating the non-random nature of circRNA alterations. Most circRNAs detected in blood were also present in the brain. Therefore, circRNAs may play potential functional roles after IS, and circRNAs in the blood may have the same origin as circRNAs in the ischemic brain [200]. Peripheral inflammatory cells possibly pass the BBB through chemotaxis and reach the brain tissue at the ischemic penumbra to exert an inflammatory response.

4.1.2 Peripheral Blood Immune Cells Participate in the Inflammatory Response after Stroke through the Function of circRNAs

Recently, with advances in clinical trials, the relationship between circRNAs and IS has become increasingly evident. Clinical trial results are highly consistent with those of animal trials.

In a study by Dong Z et al. [201] in 2020 examining circRNA expression in stroke, the expression profile of 164532 circRNAs was generated in peripheral blood mononuclear cells (PBMCs) from patients with acute ischemic stroke (AIS) and controls using high-throughput sequencing technology. Among these, 521 circRNAs were differentially expressed, with 373 upregulated and 148 downregulated. The study demonstrated significant differences in circRNAs in PBMCs between patients with AIS and controls. Additionally, this study indicated that circRNAs participated in the inflammatory pathways (such as the NF-κB, the TNF, and the chemokine signaling pathways) by acting as miRNA sponges [201]. However, the specific pathways of circRNA action were not studied in more detail in this paper. Table 2 (Ref. [202, 203, 204, 205, 206, 207, 208, 209, 210, 211, 212, 213, 214, 215, 216, 217, 218, 219]) outlines the role of circRNAs in modulating inflammation.

Table 2.The evidence of circRNAs regulatory roles in the pathophysiological changes of ischemic stroke.
Pathological process Cell line/treatment Animal/disease model CircRNAs and expression Regulating axis Overexpression role References
Angiogenesis HBMECs / CircFUNDC1 UP miR-375/PTEN Inhibit angiogenesis [202]
OGD-treated Detrimental
HBMECs/ tMCAO CircPDS5B UP hnRNPL/VEGFA Inhibit angiogenesis [203]
OGD-treated Mouse Detrimental
/ tMCAO Circ-DLGAP4 DOWN miR-143/hectord1/EndoMT Decrease BBB damage [204]
Mouse Protective
Apoptosis HT22 tMCAO CircCDC14A UP miR-23a-3p/CXCL12 Induce apoptosis [205, 206]
OGD/R Mice Detrimental
astrocyte tMCAO Circ-CELF1 UP DDX54/NFAT5 Induce apoptosis [207]
OGD/R Mouse Detrimental
HT22 MCAO Circ-HECTD1 DOWN miR-125b-5p/GDF11 Inhibit apoptosis [208]
OGD Mice Protective
Inflammation / tMCAO CircPUM1 UP miR-340-5p/DDX5/NF-κB Inhibit inflammation [209]
Mouse Protective
microglia / CircPTK2 UP miR-29b-SOCS-1-JAK2/STAT3-IL-1β Induce inflammation [210]
OGD Detrimental
microglia tMCAO Circ-0000831 UP miR-16-5p/AdipoR2 Induce inflammation [211]
OGD Mice Detrimental
HCN-2 / Circ-0007290 UP miR-496/PDCD4 Induce inflammation [212]
OGD Detrimental
HBMVECs / Circ-Memo1 UP miR-17-5p/SOS1 Induce inflammation [213]
H/R Detrimental
HBMECs / Circ-0006459 UP miR-940/FOXJ2 Induce inflammation [214]
OGD Detrimental
HBMECs / Circ-0000566 UP miR-18a-5p/ACVR2B Induce inflammation [215]
OGD/R Detrimental
SK-N-SH / Circ-0000647 UP miR-126-5p/TRAF3 Induce inflammation [216]
OGD/R Detrimental
SK-N-SH tMCAO Circ-0101874 UP miR-335-5p/PDE4D Induce inflammation [217]
OGD Mice Detrimental
HT22 MCAO Circ-HECTD1 UP miR-133b/TRAF3 Induce inflammation [208]
OGD/R Mouse Detrimental
HCN-2 / Circ-DLGAP4 DOWN miR-503-3p/NEGR1 Inhibit inflammation [218]
OGD Protective
astrocyte MCAO Circ-CTNNB1 DOWN miR-96-5p/SRB1 Inhibit inflammation [219]
OGD/R Mouse Protective

HBMECs, human brain microvascular endothelial cells; OGD, oxygen-glucose deprivation; HT22, hippocampal neuronal cell line; HCN-2, cortical neuron cell line; HBMVECs, human brain microvascular endothelial cells; H/R, hypoxia/reoxygenation; SK-N-SH, human neuroblastoma cell line; tMCAO, transient middle cerebral artery occlusion; PTEN, phosphatase and tensin homolog; hnRNPL/VEGFA, heterogenous nuclear ribonucleoprotein L /vascular endothelial growth factor-A; EndoMT, endothelial-mesenchymal transition; CXCL12, C-X-C motif chemokine ligand-12; DDX54, DEAD-box helicase 54; NFAT5, nuclear factor of activated T cell 5; GDF11, growth differentiation factor 11; NF-κB, nuclear factor kappa-B; miR-29b-SOCS-1-JAK2/STAT3-IL-1β, suppressor of cytokine signaling 1-Janus kinase2/signal transducer and activator of transcription 3-interleukin-1β; PDCD4, programmed cell death protein 4; SOS1, Son of Sevenless 1; FOXJ2, forkhead box J2; ACVR2B, activin receptor type 2B; TRAF3, TNF receptor associated factor 3; PDE4D, phosphodiesterase 4D; NEGR1, neuronal growth regulator 1; SRB1, scavenger receptor class B type 1; OGD/R, oxygen-glucose deprivation/reoxygenation.

4.1.2.1 CircFUNDC1, circPDS5B, and circCDC14A

Another study exploring the role of peripheral blood circRNAs in IS found that among the 10798 circRNAs analyzed in human blood, 68 downregulated and 10 upregulated circRNAs differed significantly between groups. Moreover, hsa_circ_0007290 (circFUNDC1), hsa_circ_0004494 (circPDS5B), and hsa_circ_0000097 (circCDC14A) levels were positively correlated with cerebral infarct volume. circPDS5B levels showed significantly higher expression in peripheral blood lymphocytes, granulocytes, and plasma in patients with stroke, whereas circCDC14A was significantly upregulated in granulocytes [220]. Bai et al. [202] demonstrated the role of circFUNDC1 in human IS and oxygen-glucose deprivation (OGD)-treated cell models. In both model types, circFUNDC1 was upregulated, resulting in low miR-375 levels. CircFUNDC1 knockdown and miR-375 overexpression enhanced the angiogenesis ability of human brain microvascular endothelial cells (HBMECs) by regulating phosphatase and tensin homolog (PTEN) expression. Previous research has shown that PTEN can regulate BBB permeability and neurogenesis [219, 221]. Jiang et al. [203] found that circPDSB5 was also upregulated in the serum of patients with IS and tMCAO mice. Moreover, circPDS5B knockdown alleviated cell necrosis and provided cerebrovascular protection via the heterogenous nuclear ribonucleoprotein L/vascular endothelial growth factor-A (hnRNPL/VEGFA) pathway. Zuo et al. [205] noted the high expression of circCDC14A in peripheral plasma, neutrophils, pericerebral infarction cortex, and astrocytes. Reducing circCDC14A levels in peripheral blood immune cells alleviated astrocyte activation in the periinfarction cortex, thus alleviating brain injury in AIS. This may be because peripheral neutrophils infiltrated ischemic brain tissue through the BBB. Additionally, Huo et al. [206] observed that circCDC14A positively modulated C-X-C motif chemokine ligand-12 (CXCL12), inducing neuronal damage and apoptosis via miR-23a-3p in tMCAO mice.

4.1.2.2 CircDLGAP4

In 2019, Zhu et al. [222] identified a correlation between has_circ_0060180 (circDLGAP4) expression and the inflammatory cascade level in patients with AIS. CircDLGAP4 was down-regulated in PBMCs in patients with AIS and negatively associated with miR-143 expression. circDLGAP4 downregulation and miR-143 upregulation increased serum inflammatory cytokines including IL-6, IL-8, IL-22, and TNF-α. However, the relevance of miR-143 in acute stroke remains unclear and may be related to BBB breakdown and PBMC activation [222]. Similarly, Bai et al. [204] observed abnormal circDLGAP4 expression in PBMCs and found that circDLGAP4 acts as a miR-143 sponge to reduce miR-143 level, causing downregulation of homologous to the E6-AP C-terminal domain E3 ubiquitin protein ligase 1 (HECTD1) expression in the ischemic brain. CircDLGAP4 overexpression significantly inhibited endothelial-mesenchymal transition to decrease BBB damage and infarct areas by modulating the expression of tight junction protein and mesenchymal cell marker in the tMCAO mouse stroke model. In another cell experiment, Qiu et al. [218] confirmed that circDLGAP4 functions as a miR-503-3p sponge to elevate neuronal growth regulator 1 (NEGR1) expression and influence the level of the proinflammatory cytokine, thereby attenuating inflammatory reaction and neuronal impairments during IS.

4.2 CircRNAs Affect Microglia and Astrocytes through the Inflammatory Processes
4.2.1 CircPUM1

Hu et al. [209] observed significant upregulation of hsa_circ_0000043 (circPUM1) and DEAD-box helicase 5 (DDX5) and notable miR-340-5p downregulation in a tMCAO mouse model [209]. Abnormally low circPUM1 expression promoted the secretion of inflammatory factors and induced apoptosis by activating microglia and astrocytes through the miR-340-5p/DDX5/ NF-κB pathway, ultimately causing ischemic injury in circPUM1 knockout mice. circPUM1 overexpression acted as miR-340-3p sponge, upregulating DDX5 and reducing the content of pro-inflammatory factors (IL-1β, IL-6, TNF-α), and the NF-κB (p-p65, cleaved caspase-3, and Bax) expression [209, 223]. Moreover, miR-340 has been implicated in various diseases, such as multiple myeloma, pulmonary sarcoidosis, heart failure, lung adenocarcinoma, colorectal cancer, and osteoporosis [224, 225, 226, 227, 228, 229].

4.2.2 CircPTK2

In 2019, Wang et al. [210] found that hsa_circ_0005273 (circPTK2) regulates neuronal apoptosis by activating microglia via miR-29b-suppressor of cytokine signaling 1-Janus kinase 2/signal transducer and activator of transcription 3-interleukin-1β (SOCS-1-JAK2/STAT3-IL-1β) signaling. circPTK2 inhibits miR-29b activity by acting as a miRNA sponge, thus downregulating the suppressor of cytokine signaling 1 (SOCS-1) in microglia. SOCS-1 targets Janus kinase 2 (JAK2) and inhibits the activation of the JAK2/STAT3 pathway [230, 231]. STAT3 tyrosine phosphorylation induced IL-1β and IL-6 production in the inflammation response [232]. Therefore, circPTK2 promotes neuronal inflammation and apoptosis by regulating microglia [210].

4.2.3 Circ-0000831

It has been reported that circ-0000831 mitigated inflammatory and apoptotic responses in MCAO mice models, and circ-0000831 was upregulated in microglia after OGD. circ-0000831 acted as a sponge for miR-16-5p, causing its downregulation. Adiponectin receptor 2 (AdipoR2), a target for miR-16-5p, was upregulated and increased peroxisome proliferator-activated receptor γ (PPARγ) expression in microglia in animal models [211]. AdipoR2 has been reported to activate PPARγ pathways to modify inflammatory processes, and PPARγ overexpression alleviated the inflammation response [233, 234]. Moreover, AdipoR2 activated the PPARα pathway to inhibit inflammation [235]. Therefore, circ-0000831 can alleviate neuroinflammation and cell apoptosis in MCAO mice via the miR-15-5p-AdipoR2-PPARγ pathway [211].

4.2.4 CircCELF1

Li et al. [207] observed that highly expressed circCELF1 in oxygen-glucose deprivation/reoxygenation (OGD/R) astrocytes could exacerbate astrocyte damage, and its expression is associated with inflammatory response. CircRNA hsa_circ_0000304 (circCELF1) knockdown decreased the expression of proinflammatory cytokines in astrocytes after OGD/R. By binding to the RNA binding protein DEAD-box helicase 54 (DDX54), circCELF1 enhances the stability of the nuclear factor of activated T cell 5 (NFAT5) mRNA. In another study, increased NFAT5 expression induced neuroinflammation after IS [236]. NFAT5 overexpression can restore the decline in inflammatory factor levels caused by circCELF1 knockdown [207].

4.2.5 CircCTNNB1

Recently, circRNA catenin beta 1 (circCTNNB1) was found to play a role in the inflammatory response in astrocytes. CircCTNNB1 and IL-10 levels were downregulated, whereas the expression of proinflammatory factors, including IL-1β, IL-6, and TNF-α, was upregulated when mouse astrocytes treated with OGD/R treatment. By sponging miR-96-5p, circCTNNB1 upregulation increased scavenger receptor class B type 1 (SRB1) levels to protect against brain injury [237]. SRB1 is involved in cerebrovascular diseases. However, there is no evidence proving the correlation between SRB1 and the BBB, necessitating further research.

4.3 CircRNAs Affect other Cells through the Inflammatory Processes
4.3.1 CircMemo1

Vascular endothelial cells, an important part of the BBB, were widely implicated in the inflammatory response after IS. Ren et al. [213] found that circ-Memo1 and Son of Sevenless 1 (SOS1) levels were highly expressed; however, miR-17-5p expression was low in the peripheral blood of patients with IS. After circMemo1 knockdown, the expression of proinflammatory cytokines was low in the human brain microvascular endothelial cell (HBMVEC) model. CircMemo1 can directly bind with miR-17-5p, negatively regulating miR-17-5p levels, which could inhibit SOS1 expression. Cerebral ischemia stimulates SOS1 activity and disrupts Extracellular Signal-Regulated Kinase 1/2 (ERK1/2) phosphorylation downstream of SOS1 [238]. Moreover, SOS1 can activate the NF-κB pathway by regulating epidermal growth factor (EGF) levels [239]. Inhibiting the ERK1/2 pathway could enhance NF-κB phosphorylation and activity and reduce brain injury. However, ERK1/2 pathway activation promotes inflammation after cerebral ischemia [240, 241].

4.3.2 Circ-0006459

Li et al. [214] investigated the inflammatory response of circ-0006459 in OGD-induced HBMEC cell models. By sponging miR-940, circ-0006459 overexpression promoted proinflammatory factor levels (including IL-1β, IL-8, IL-18, and TNF-α) and induced forkhead box J2 (FOXJ2) upregulation in HBMEC injury after OGD. FOXJ2 has DNA-binding specificity and is involved in neuroinflammatory responses as a transcription factor [242, 243]. This pathway inhibits HBMEC proliferation and promotes inflammatory responses.

4.3.3 Circ-0000566

In a HBMECs model article, it has been discovered that circ-0000566 was highly expressed after OGD [244]. Another study found that circ-0000566 was upregulated after OGD/R, while miR-18a-5p was significantly downregulated. Moreover, Circ-0000566 acted as a miR-18a-5p sponge, contributing to HBMEC injury and proinflammatory cytokine release [215]. Activin receptor type IIB (ACVR2B) directly targeted miR-18a-5p, causing its upregulation and involvement in the smad2/c-jun axis in neuronal damage after cerebral ischemia in mice [245].

4.3.4 Circ-0000647

In 2022, Dai et al. [216] used human neuroblastoma cell line (SK-N-SH) cells to simulate the cerebral ischemia environment and found elevated circ-0000647 levels, with circ-0000647 acting as a miRNA sponge to bind miR-126-5p and decrease miR-126-5p expression [216]. Circ-0000647 upregulation and miR-126-5pc downregulation reduced the expression of TNF-α and IL-6. Mechanistically, TNF receptor-associated factor 3 directly targeted miR-126-5p and reduced TNF-α and IL-6 expressions. Therefore, circ-0000647 can promote inflammation via the miR-126-5p-TRAF3 (TNF receptor associated factor 3) pathway [216].

4.3.5 Circ-0101874

In this review, the involvement of circRNA in IS by influencing inflammatory factors has been summarized. Pei et al. [217] observed that circ-0101874 was upregulated in both MCAO mice and OGD-induced cell models. Circ-0101874 knockdown suppressed TNF-α and IL-6 expression in OGD cells. Therefore, miR-335-5p acts as the target of circ-0101864 and deactivates the gene called phosphodiesterase 4D to abate the inflammatory response.

4.3.6 CircHECTD1

The role of circular RNA HECT domain E3 ubiquitin-protein ligase 1 (circHECTD1) in mouse MCAO and OGD cell models has been found in 2021. CircHECTD1 expression was upregulated, and its interaction with miR-133b reduced its expression. TNF receptor-associated factor 3, the target for miR-133b, was highly expressed and promoted the expression of inflammatory cytokines such as NF-κB [208].

4.3.7 CircUCK2

In another cell model, hsa_circ_001,357 (circUCK2) inhibited miR-125b-5p expression by acting as a miRNA sponge and increased the expression of growth differentiation factor 11 (GDF11) [246]. GDF11 is a member of the TGF-β superfamily and binds with TGF-β type I and II receptors to activate the mothers against the decapentaplegic homolog 3 (Smad3) signaling pathway [247]. Smad3 overexpression inhibits inflammation and reduces cell damage after ischemic cerebral infarction [248].

4.3.8 Circ-0007290

Wang et al. [212] identified the inflammatory effects of circ-0007290 expression in both in vitro and in vivo models. Circ-0007290 expression was elevated, and knockdown circ-0007290 knockdown reduced the inflammatory response after OGD. Circ-0007290 acted as a sponge to mitigate miR-496 inhibition of its target, programmed cell death protein 4. Moreover, circ-0007290 inhibition, miR-496 overexpression, or programmed cell death protein 4 upregulation induced proinflammatory factor (such as TNF-α and IL-1β) release.

5. Discussion

IS, a condition causing cerebral tissue death and neuronal damage, still has a high morbidity and mortality rate [249]. Inflammation is an important pathological reaction in AIS, and circRNAs may play critical roles in it [250, 251]. CircRNAs are not only potential novel biomarkers but also vital for therapeutic targets for the diagnosis and treatment of IS [252]. After IS, circRNAs are differentially expressed in neurons and inflammatory cells affecting the release of inflammatory mediators. Among them, circRNAs mainly act as sponges to combined with miRNAs so that inhibit the effect of miRNA and increase the expression level of target genes. However, only once study has focused on the function of forming circRNP complexes. Furthermore, circRNAs exhibit diverse functions through multiple pathways. As research progresses, the distinct mechanisms of regulation can be better elucidated, thereby enhancing our understanding of the inflammatory regulatory network involving circRNAs.

Currently, the potential of utilizing circRNAs as targets for early diagnosis and post-treatment analysis is yet to be determined. However, it can be concluded that circRNA is widely involved in the activation of various inflammatory cells, the secretion of inflammatory mediators, and the destruction of the BBB. By modulating circRNAs, it is possible to change the extent of inflammatory response. The regulatory mechanisms of inflammation are incredibly intricate, and the clinical application still needs to face many challenges. Moreover, because of the different etiology of the IS subtypes (such as atherothrombotic IS, cardiac cerebral embolism and so on), it’s difficult to explain the inflammatory response in the development of IS [253]. However, the ultimate result of various etiological subtypes is to lead to vascular occlusion, and the inflammatory factors and inflammatory cells involved in the process are the same. Therefore, it is meaningful to use the cell model of oxygen-glucose deprivation and the mouse model of MCAO to study the inflammatory response after IS.

Due to the complexity of etiological subtypes and the difficulty of model building, there are few studies on the involvement of circRNA in etiology. It is great significance to explore the role of circRNA in the pathogenesis of different subtypes of IS. With the advancement of technology, the establishment of more appropriate and realistic models should lead to more perfect results.

6. Conclusion

In this review, we described the inflammatory response after IS and summerized the roles of circRNAs in IS. In all, the immune response in IS is quite complex and circRNAs show dual roles to regulate the neuro-inflammation. Restricting the dentrimental neuroinflammation might be helpful to decrease ischemic volume. CircRNA light a new potential treatment target for IS. In order to determine the effective therapeutic strategy, is important to investigate the mechanisms between circRNA and other inflammatory factors.

Abbreviations

BBB, Blood-brain barrier; circRNA, circular RNA; DAMPs, damage associated molecular patterns; HMGB1, high-mobility group box 1 protein; TLRs, Toll-like receptors; TNF-α, tumor necrosis factor-α; IL, interleukin; TGF-β, transforming growth factor-β; MIP-1, macrophage inflammatory protein-1; MCP-1, monocyte chemoattractant protein-1; IP-10 (CXCL10), chemokines interferon-γ-inducible protein 10; MMP, matrix metalloproteinase; IFN-γ, interferon-γ; ICAM-1, intercellular cell adhesion molecule-1; VCAM-1, vascular endothelial cell adhesion molecule-1; MIF, macrophage migration inhibitory factor; NK cells, Natural killer cells; Treg cells, regulatory T cells; NMDARs, N-Methyl-D-aspartic acid receptors; RNA Pol II, RNA polymerase II; TFs, transcription factors; RBPs, RNA binding proteins; ecircRNAs, exonic circRNAs; ciRNAs, circular intronic RNAs; EIciRNAs, exon-intron circRNAs; tricRNAs, tRNA intronic circular RNAs; MCAO, middle cerebral artery occlusion; PBICs, Peripheral blood immune cells; AIS, acute ischemic stroke; PBMCs, peripheral blood mononuclear cells; NF-κB, nuclear factor kappa-B; tMCAO, transient middle cerebral artery occlusion; PTEN, phosphatase and tensin homolog; CXCL12, C-X-C motif chemokine ligand-12; HECTD1, homologous to the E6-AP C-terminal domain E3 ubiquitin protein ligase 1; NEGR1, neuronal growth regulator 1; DDX5, DEAD-box helicase 5; PPARγ, peroxisome proliferator-activated receptor γ; NFAT-5, nuclear factor of activated T cells 5; FOXJ2, forkhead box J2; ACVR2B, Activin receptor type IIB; TRAF3, TNF receptor associated factor 3; PDE4D, phosphodiesterase 4D; GDF11, growth differentiation factor 11; Smad3, mothers against decapentaplegic homolog 3; OGD/R, oxygen-glucose deprivation/reoxygenation; circRNP, circRNA-protein.

Author Contributions

LJ, AWS and CGY proposed this research direction. XG, BC, YYZ, XQL, HQS and ZZ conceptualized, wrote, and edited the manuscript. XG, YJZ, KQL, WQX, LFL, JYH and BBG designed the figures and tables, drafted and edited the manuscript. LFL, KQL, JYH, BBG and CGY helped in revision and edited the manuscript. All authors contributed to editorial changes in the manuscript. All authors read and approved the final manuscript. All authors have participated sufficiently in the work and agreed to be accountable for all aspects of the work.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

Not applicable.

Funding

This work is supported by the Hangzhou Science and Technology Bureau (2021WJCY009), Hangzhou Medical and Health Technology Project (0020190131), Zhejiang Provincial Public Welfare (LGF18H090017) and Zhejiang Provincial Natural Science Foundation of China (LQ21H090005).

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
GBD 2019 Stroke Collaborators. Global, regional, and national burden of stroke and its risk factors, 1990-2019: a systematic analysis for the Global Burden of Disease Study 2019. The Lancet. Neurology. 2021; 20: 795–820.
[2]
Randolph SA. Ischemic Stroke. Workplace Health & Safety. 2016; 64: 444.
[3]
Rajkovic O, Potjewyd G, Pinteaux E. Regenerative Medicine Therapies for Targeting Neuroinflammation After Stroke. Frontiers in Neurology. 2018; 9: 734.
[4]
Barone FC, Feuerstein GZ. Inflammatory mediators and stroke: new opportunities for novel therapeutics. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 1999; 19: 819–834.
[5]
Samson Y, Lapergue B, Hosseini H. Inflammation and ischaemic stroke: current status and future perspectives. Revue Neurologique. 2005; 161: 1177–1182.
[6]
Chamorro A, Hallenbeck J. The harms and benefits of inflammatory and immune responses in vascular disease. Stroke. 2006; 37: 291–293.
[7]
Muir KW, Tyrrell P, Sattar N, Warburton E. Inflammation and ischaemic stroke. Current Opinion in Neurology. 2007; 20: 334–342.
[8]
Zhao SC, Ma LS, Chu ZH, Xu H, Wu WQ, Liu F. Regulation of microglial activation in stroke. Acta Pharmacologica Sinica. 2017; 38: 445–458.
[9]
Tuttolomondo A, Di Raimondo D, di Sciacca R, Pinto A, Licata G. Inflammatory cytokines in acute ischemic stroke. Current Pharmaceutical Design. 2008; 14: 3574–3589.
[10]
Memczak S, Jens M, Elefsinioti A, Torti F, Krueger J, Rybak A, et al. Circular RNAs are a large class of animal RNAs with regulatory potency. Nature. 2013; 495: 333–338.
[11]
Siracusa C, Vono N, Morano MB, Sabatino J, Leo I, Eyileten C, et al. Clinical Application of Circular RNAs as Biomarkers in Acute Ischemic Stroke. Journal of Personalized Medicine. 2023; 13: 839.
[12]
Yang L, Han B, Zhang Z, Wang S, Bai Y, Zhang Y, et al. Extracellular Vesicle-Mediated Delivery of Circular RNA SCMH1 Promotes Functional Recovery in Rodent and Nonhuman Primate Ischemic Stroke Models. Circulation. 2020; 142: 556–574.
[13]
Conti A, Giovannini L, Mandel VD, Odorici G, Lasagni C, Bigi L, et al. Chronic kidney disease in psoriasis: a cohort study. Journal of the German Society of Dermatology: JDDG. 2020; 18: 438–445.
[14]
Zhou D, Huang Z, Zhu X, Hong T, Zhao Y. Circular RNA 0025984 Ameliorates Ischemic Stroke Injury and Protects Astrocytes Through miR-143-3p/TET1/ORP150 Pathway. Molecular Neurobiology. 2021; 58: 5937–5953.
[15]
Liu Y, Li Y, Zang J, Zhang T, Li Y, Tan Z, et al. CircOGDH Is a Penumbra Biomarker and Therapeutic Target in Acute Ischemic Stroke. Circulation Research. 2022; 130: 907–924.
[16]
Yang Z, Huang C, Wen X, Liu W, Huang X, Li Y, et al. Circular RNA circ-FoxO3 attenuates blood-brain barrier damage by inducing autophagy during ischemia/reperfusion. Molecular Therapy: the Journal of the American Society of Gene Therapy. 2022; 30: 1275–1287.
[17]
Han B, Zhang Y, Zhang Y, Bai Y, Chen X, Huang R, et al. Novel insight into circular RNA HECTD1 in astrocyte activation via autophagy by targeting MIR142-TIPARP: implications for cerebral ischemic stroke. Autophagy. 2018; 14: 1164–1184.
[18]
Liu X, Wang Q, Zhao J, Chang H, Zhu R. Inflammation-Related circRNA Polymorphism and Ischemic Stroke Prognosis. Journal of Molecular Neuroscience: MN. 2021; 71: 2126–2133.
[19]
Moskowitz MA, Lo EH, Iadecola C. The science of stroke: mechanisms in search of treatments. Neuron. 2010; 67: 181–198.
[20]
Khoshnam SE, Winlow W, Farzaneh M, Farbood Y, Moghaddam HF. Pathogenic mechanisms following ischemic stroke. Neurological Sciences: Official Journal of the Italian Neurological Society and of the Italian Society of Clinical Neurophysiology. 2017; 38: 1167–1186.
[21]
McColl BW, Allan SM, Rothwell NJ. Systemic infection, inflammation and acute ischemic stroke. Neuroscience. 2009; 158: 1049–1061.
[22]
Liu T, Clark RK, McDonnell PC, Young PR, White RF, Barone FC, et al. Tumor necrosis factor-alpha expression in ischemic neurons. Stroke. 1994; 25: 1481–1488.
[23]
Ferrarese C, Mascarucci P, Zoia C, Cavarretta R, Frigo M, Begni B, et al. Increased cytokine release from peripheral blood cells after acute stroke. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 1999; 19: 1004–1009.
[24]
Kono H, Rock KL. How dying cells alert the immune system to danger. Nature Reviews. Immunology. 2008; 8: 279–289.
[25]
Schuhmann MK, Kollikowski AM, März AG, Bieber M, Pham M, Stoll G. Danger-associated molecular patterns are locally released during occlusion in hyper-acute stroke. Brain, Behavior, & Immunity - Health. 2021; 15: 100270.
[26]
Singh V, Roth S, Veltkamp R, Liesz A. HMGB1 as a Key Mediator of Immune Mechanisms in Ischemic Stroke. Antioxidants & Redox Signaling. 2016; 24: 635–651.
[27]
Rubartelli A, Lotze MT. Inside, outside, upside down: damage-associated molecular-pattern molecules (DAMPs) and redox. Trends in Immunology. 2007; 28: 429–436.
[28]
Petrovic-Djergovic D, Goonewardena SN, Pinsky DJ. Inflammatory Disequilibrium in Stroke. Circulation Research. 2016; 119: 142–158.
[29]
Benakis C, Garcia-Bonilla L, Iadecola C, Anrather J. The role of microglia and myeloid immune cells in acute cerebral ischemia. Frontiers in Cellular Neuroscience. 2015; 8: 461.
[30]
Marsh BJ, Williams-Karnesky RL, Stenzel-Poore MP. Toll-like receptor signaling in endogenous neuroprotection and stroke. Neuroscience. 2009; 158: 1007–1020.
[31]
Gill R, Tsung A, Billiar T. Linking oxidative stress to inflammation: Toll-like receptors. Free Radical Biology & Medicine. 2010; 48: 1121–1132.
[32]
Cătălin B, Cupido A, Iancău M, Albu CV, Kirchhoff F. Microglia: first responders in the central nervous system. Romanian Journal of Morphology and Embryology. 2013; 54: 467–472.
[33]
Hoehn BD, Palmer TD, Steinberg GK. Neurogenesis in rats after focal cerebral ischemia is enhanced by indomethacin. Stroke. 2005; 36: 2718–2724.
[34]
Salter MW, Stevens B. Microglia emerge as central players in brain disease. Nature Medicine. 2017; 23: 1018–1027.
[35]
Michelucci A, Heurtaux T, Grandbarbe L, Morga E, Heuschling P. Characterization of the microglial phenotype under specific pro-inflammatory and anti-inflammatory conditions: Effects of oligomeric and fibrillar amyloid-beta. Journal of Neuroimmunology. 2009; 210: 3–12.
[36]
Orihuela R, McPherson CA, Harry GJ. Microglial M1/M2 polarization and metabolic states. British Journal of Pharmacology. 2016; 173: 649–665.
[37]
Gogoleva VS, Drutskaya MS, Atretkhany KSN. The Role of Microglia in the Homeostasis of the Central Nervous System and Neuroinflammation. Molekuliarnaia Biologiia. 2019; 53: 790–798.
[38]
Lee YB, Nagai A, Kim SU. Cytokines, chemokines, and cytokine receptors in human microglia. Journal of Neuroscience Research. 2002; 69: 94–103.
[39]
Zhu J, Cao D, Guo C, Liu M, Tao Y, Zhou J, et al. Berberine Facilitates Angiogenesis Against Ischemic Stroke Through Modulating Microglial Polarization via AMPK Signaling. Cellular and Molecular Neurobiology. 2019; 39: 751–768.
[40]
Patel AR, Ritzel R, McCullough LD, Liu F. Microglia and ischemic stroke: a double-edged sword. International Journal of Physiology, Pathophysiology and Pharmacology. 2013; 5: 73–90.
[41]
Allen NJ, Eroglu C. Cell Biology of Astrocyte-Synapse Interactions. Neuron. 2017; 96: 697–708.
[42]
Azevedo FAC, Carvalho LRB, Grinberg LT, Farfel JM, Ferretti REL, Leite REP, et al. Equal numbers of neuronal and nonneuronal cells make the human brain an isometrically scaled-up primate brain. The Journal of Comparative Neurology. 2009; 513: 532–541.
[43]
Lee HG, Wheeler MA, Quintana FJ. Function and therapeutic value of astrocytes in neurological diseases. Nature Reviews. Drug Discovery. 2022; 21: 339–358.
[44]
Liddelow SA, Guttenplan KA, Clarke LE, Bennett FC, Bohlen CJ, Schirmer L, et al. Neurotoxic reactive astrocytes are induced by activated microglia. Nature. 2017; 541: 481–487.
[45]
Barreto G, White RE, Ouyang Y, Xu L, Giffard RG. Astrocytes: targets for neuroprotection in stroke. Central Nervous System Agents in Medicinal Chemistry. 2011; 11: 164–173.
[46]
Dimitrijevic OB, Stamatovic SM, Keep RF, Andjelkovic AV. Effects of the chemokine CCL2 on blood-brain barrier permeability during ischemia-reperfusion injury. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 2006; 26: 797–810.
[47]
Lau LT, Yu AC. Astrocytes produce and release interleukin-1, interleukin-6, tumor necrosis factor alpha and interferon-gamma following traumatic and metabolic injury. Journal of Neurotrauma. 2001; 18: 351–359.
[48]
Ronaldson PT, Davis TP. Blood-brain barrier integrity and glial support: mechanisms that can be targeted for novel therapeutic approaches in stroke. Current Pharmaceutical Design. 2012; 18: 3624–3644.
[49]
Yoshimoto T, Houkin K, Tada M, Abe H. Induction of cytokines, chemokines and adhesion molecule mRNA in a rat forebrain reperfusion model. Acta Neuropathologica. 1997; 93: 154–158.
[50]
Kaplanski G, Farnarier C, Kaplanski S, Porat R, Shapiro L, Bongrand P, et al. Interleukin-1 induces interleukin-8 secretion from endothelial cells by a juxtacrine mechanism. Blood. 1994; 84: 4242–4248.
[51]
Weller A, Isenmann S, Vestweber D. Cloning of the mouse endothelial selectins. Expression of both E- and P-selectin is inducible by tumor necrosis factor alpha. The Journal of Biological Chemistry. 1992; 267: 15176–15183.
[52]
Carvalho-Tavares J, Hickey MJ, Hutchison J, Michaud J, Sutcliffe IT, Kubes P. A role for platelets and endothelial selectins in tumor necrosis factor-alpha-induced leukocyte recruitment in the brain microvasculature. Circulation Research. 2000; 87: 1141–1148.
[53]
Henninger DD, Panés J, Eppihimer M, Russell J, Gerritsen M, Anderson DC, et al. Cytokine-induced VCAM-1 and ICAM-1 expression in different organs of the mouse. Journal of Immunology (Baltimore, Md.: 1950). 1997; 158: 1825–1832.
[54]
Bui TA, Jickling GC, Winship IR. Neutrophil dynamics and inflammaging in acute ischemic stroke: A transcriptomic review. Frontiers in Aging Neuroscience. 2022; 14: 1041333.
[55]
Huang J, Upadhyay UM, Tamargo RJ. Inflammation in stroke and focal cerebral ischemia. Surgical Neurology. 2006; 66: 232–245.
[56]
Perez-de-Puig I, Miró-Mur F, Ferrer-Ferrer M, Gelpi E, Pedragosa J, Justicia C, et al. Neutrophil recruitment to the brain in mouse and human ischemic stroke. Acta Neuropathologica. 2015; 129: 239–257.
[57]
Jayaraj RL, Azimullah S, Beiram R, Jalal FY, Rosenberg GA. Neuroinflammation: friend and foe for ischemic stroke. Journal of Neuroinflammation. 2019; 16: 142.
[58]
Gelderblom M, Leypoldt F, Steinbach K, Behrens D, Choe CU, Siler DA, et al. Temporal and spatial dynamics of cerebral immune cell accumulation in stroke. Stroke. 2009; 40: 1849–1857.
[59]
Geissmann F, Manz MG, Jung S, Sieweke MH, Merad M, Ley K. Development of monocytes, macrophages, and dendritic cells. Science (New York, N.Y.). 2010; 327: 656–661.
[60]
Garcia JH, Liu KF, Yoshida Y, Lian J, Chen S, del Zoppo GJ. Influx of leukocytes and platelets in an evolving brain infarct (Wistar rat). The American Journal of Pathology. 1994; 144: 188–199.
[61]
Urra X, Cervera A, Obach V, Climent N, Planas AM, Chamorro A. Monocytes are major players in the prognosis and risk of infection after acute stroke. Stroke. 2009; 40: 1262–1268.
[62]
Gan Y, Liu Q, Wu W, Yin JX, Bai XF, Shen R, et al. Ischemic neurons recruit natural killer cells that accelerate brain infarction. Proceedings of the National Academy of Sciences of the United States of America. 2014; 111: 2704–2709.
[63]
Ao LY, Yan YY, Zhou L, Li CY, Li WT, Fang WR, et al. Immune Cells After Ischemic Stroke Onset: Roles, Migration, and Target Intervention. Journal of Molecular Neuroscience: MN. 2018; 66: 342–355.
[64]
Engelhardt B, Sorokin L. The blood-brain and the blood-cerebrospinal fluid barriers: function and dysfunction. Seminars in Immunopathology. 2009; 31: 497–511.
[65]
Chamorro Á, Meisel A, Planas AM, Urra X, van de Beek D, Veltkamp R. The immunology of acute stroke. Nature Reviews. Neurology. 2012; 8: 401–410.
[66]
Abbas AK, Murphy KM, Sher A. Functional diversity of helper T lymphocytes. Nature. 1996; 383: 787–793.
[67]
Martin B, Hirota K, Cua DJ, Stockinger B, Veldhoen M. Interleukin-17-producing gammadelta T cells selectively expand in response to pathogen products and environmental signals. Immunity. 2009; 31: 321–330.
[68]
Shichita T, Sugiyama Y, Ooboshi H, Sugimori H, Nakagawa R, Takada I, et al. Pivotal role of cerebral interleukin-17-producing gammadeltaT cells in the delayed phase of ischemic brain injury. Nature Medicine. 2009; 15: 946–950.
[69]
Yilmaz G, Arumugam TV, Stokes KY, Granger DN. Role of T lymphocytes and interferon-gamma in ischemic stroke. Circulation. 2006; 113: 2105–2112.
[70]
Liesz A, Zhou W, Mracskó É, Karcher S, Bauer H, Schwarting S, et al. Inhibition of lymphocyte trafficking shields the brain against deleterious neuroinflammation after stroke. Brain: a Journal of Neurology. 2011; 134: 704–720.
[71]
Offner H, Subramanian S, Parker SM, Wang C, Afentoulis ME, Lewis A, et al. Splenic atrophy in experimental stroke is accompanied by increased regulatory T cells and circulating macrophages. Journal of Immunology (Baltimore, Md.: 1950). 2006; 176: 6523–6531.
[72]
Liesz A, Suri-Payer E, Veltkamp C, Doerr H, Sommer C, Rivest S, et al. Regulatory T cells are key cerebroprotective immunomodulators in acute experimental stroke. Nature Medicine. 2009; 15: 192–199.
[73]
Ren X, Akiyoshi K, Dziennis S, Vandenbark AA, Herson PS, Hurn PD, et al. Regulatory B cells limit CNS inflammation and neurologic deficits in murine experimental stroke. The Journal of Neuroscience: the Official Journal of the Society for Neuroscience. 2011; 31: 8556–8563.
[74]
Zhu Y, Yang GY, Ahlemeyer B, Pang L, Che XM, Culmsee C, et al. Transforming growth factor-beta 1 increases bad phosphorylation and protects neurons against damage. The Journal of Neuroscience: the Official Journal of the Society for Neuroscience. 2002; 22: 3898–3909.
[75]
Tuttolomondo A, Pecoraro R, Pinto A. Studies of selective TNF inhibitors in the treatment of brain injury from stroke and trauma: a review of the evidence to date. Drug Design, Development and Therapy. 2014; 8: 2221–2238.
[76]
Doll DN, Rellick SL, Barr TL, Ren X, Simpkins JW. Rapid mitochondrial dysfunction mediates TNF-alpha-induced neurotoxicity. Journal of Neurochemistry. 2015; 132: 443–451.
[77]
Wilkins HM, Swerdlow RH. TNFα in cerebral ischemia: another stroke against you? Journal of Neurochemistry. 2015; 132: 369–372.
[78]
Viviani B, Bartesaghi S, Gardoni F, Vezzani A, Behrens MM, Bartfai T, et al. Interleukin-1beta enhances NMDA receptor-mediated intracellular calcium increase through activation of the Src family of kinases. The Journal of Neuroscience: the Official Journal of the Society for Neuroscience. 2003; 23: 8692–8700.
[79]
Sun GY, Hu ZY. Stimulation of phospholipase A2 expression in rat cultured astrocytes by LPS, TNF alpha and IL-1 beta. Progress in Brain Research. 1995; 105: 231–238.
[80]
Liu G, Tsuruta Y, Gao Z, Park YJ, Abraham E. Variant IL-1 receptor-associated kinase-1 mediates increased NF-kappa B activity. Journal of Immunology (Baltimore, Md.: 1950). 2007; 179: 4125–4134.
[81]
Pham LDD, Hayakawa K, Seo JH, Nguyen MN, Som AT, Lee BJ, et al. Crosstalk between oligodendrocytes and cerebral endothelium contributes to vascular remodeling after white matter injury. Glia. 2012; 60: 875–881.
[82]
Nakamura K, Shichita T. Cellular and molecular mechanisms of sterile inflammation in ischaemic stroke. Journal of Biochemistry. 2019; 165: 459–464.
[83]
Perera LP, Goldman CK, Waldmann TA. IL-15 induces the expression of chemokines and their receptors in T lymphocytes. Journal of Immunology (Baltimore, Md.: 1950). 1999; 162: 2606–2612.
[84]
Li M, Li Z, Yao Y, Jin WN, Wood K, Liu Q, et al. Astrocyte-derived interleukin-15 exacerbates ischemic brain injury via propagation of cellular immunity. Proceedings of the National Academy of Sciences of the United States of America. 2017; 114: E396–E405.
[85]
Lee GA, Lin TN, Chen CY, Mau SY, Huang WZ, Kao YC, et al. Interleukin 15 blockade protects the brain from cerebral ischemia-reperfusion injury. Brain, Behavior, and Immunity. 2018; 73: 562–570.
[86]
Zhang Q, Liao Y, Liu Z, Dai Y, Li Y, Li Y, et al. Interleukin-17 and ischaemic stroke. Immunology. 2021; 162: 179–193.
[87]
Clarkson BDS, Ling C, Shi Y, Harris MG, Rayasam A, Sun D, et al. T cell-derived interleukin (IL)-21 promotes brain injury following stroke in mice. The Journal of Experimental Medicine. 2014; 211: 595–604.
[88]
Rayasam A, Kijak JA, Kissel L, Choi YH, Kim T, Hsu M, et al. CXCL13 expressed on inflamed cerebral blood vessels recruit IL-21 producing TFH cells to damage neurons following stroke. Journal of Neuroinflammation. 2022; 19: 125.
[89]
Burnstock G. Purinergic signalling and disorders of the central nervous system. Nature Reviews. Drug Discovery. 2008; 7: 575–590.
[90]
Seifert HA, Collier LA, Chapman CB, Benkovic SA, Willing AE, Pennypacker KR. Pro-inflammatory interferon gamma signaling is directly associated with stroke induced neurodegeneration. Journal of Neuroimmune Pharmacology: the Official Journal of the Society on NeuroImmune Pharmacology. 2014; 9: 679–689.
[91]
Shi Q, Cai X, Shi G, Lv X, Yu J, Wang F. Interleukin-4 protects from chemotherapy-induced peripheral neuropathy in mice modal via the stimulation of IL-4/STAT6 signaling. Acta Cirurgica Brasileira. 2018; 33: 491–498.
[92]
Tian Y, Zhu P, Liu S, Jin Z, Li D, Zhao H, et al. IL-4-polarized BV2 microglia cells promote angiogenesis by secreting exosomes. Advances in Clinical and Experimental Medicine: Official Organ Wroclaw Medical University. 2019; 28: 421–430.
[93]
Zhu H, Hu S, Li Y, Sun Y, Xiong X, Hu X, et al. Interleukins and Ischemic Stroke. Frontiers in Immunology. 2022; 13: 828447.
[94]
Tarkowski E, Rosengren L, Blomstrand C, Wikkelsö C, Jensen C, Ekholm S, et al. Early intrathecal production of interleukin-6 predicts the size of brain lesion in stroke. Stroke. 1995; 26: 1393–1398.
[95]
Doyle KP, Simon RP, Stenzel-Poore MP. Mechanisms of ischemic brain damage. Neuropharmacology. 2008; 55: 310–318.
[96]
Gertz K, Kronenberg G, Kälin RE, Baldinger T, Werner C, Balkaya M, et al. Essential role of interleukin-6 in post-stroke angiogenesis. Brain: a Journal of Neurology. 2012; 135: 1964–1980.
[97]
de Bilbao F, Arsenijevic D, Moll T, Garcia-Gabay I, Vallet P, Langhans W, et al. In vivo over-expression of interleukin-10 increases resistance to focal brain ischemia in mice. Journal of Neurochemistry. 2009; 110: 12–22.
[98]
Sharma S, Yang B, Xi X, Grotta JC, Aronowski J, Savitz SI. IL-10 directly protects cortical neurons by activating PI-3 kinase and STAT-3 pathways. Brain Research. 2011; 1373: 189–194.
[99]
Dong Y, Hu C, Huang C, Gao J, Niu W, Wang D, et al. Interleukin-22 Plays a Protective Role by Regulating the JAK2-STAT3 Pathway to Improve Inflammation, Oxidative Stress, and Neuronal Apoptosis following Cerebral Ischemia-Reperfusion Injury. Mediators of Inflammation. 2021; 2021: 6621296.
[100]
Doyle KP, Cekanaviciute E, Mamer LE, Buckwalter MS. TGFβ signaling in the brain increases with aging and signals to astrocytes and innate immune cells in the weeks after stroke. Journal of Neuroinflammation. 2010; 7: 62.
[101]
Dobolyi A, Vincze C, Pál G, Lovas G. The neuroprotective functions of transforming growth factor beta proteins. International Journal of Molecular Sciences. 2012; 13: 8219–8258.
[102]
Hayes CA, Valcarcel-Ares MN, Ashpole NM. Preclinical and clinical evidence of IGF-1 as a prognostic marker and acute intervention with ischemic stroke. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 2021; 41: 2475–2491.
[103]
Shaheen HA, Daker LI, Abbass MM, Abd El Fattah AA. The relationship between the severity of disability and serum IL-8 in acute ischemic stroke patients. The Egyptian Journal of Neurology, Psychiatry and Neurosurgery. 2018; 54: 26.
[104]
Luscinskas FW, Kiely JM, Ding H, Obin MS, Hébert CA, Baker JB, et al. In vitro inhibitory effect of IL-8 and other chemoattractants on neutrophil-endothelial adhesive interactions. Journal of Immunology (Baltimore, Md.: 1950). 1992; 149: 2163–2171.
[105]
Brandhofer M, Hoffmann A, Blanchet X, Siminkovitch E, Rohlfing AK, El Bounkari O, et al. Heterocomplexes between the atypical chemokine MIF and the CXC-motif chemokine CXCL4L1 regulate inflammation and thrombus formation. Cellular and Molecular Life Sciences: CMLS. 2022; 79: 512.
[106]
Chen C, Chu SF, Liu DD, Zhang Z, Kong LL, Zhou X, et al. Chemokines play complex roles in cerebral ischemia. Neurochemistry International. 2018; 112: 146–158.
[107]
Zhang Y, Gao Z, Wang D, Zhang T, Sun B, Mu L, et al. Accumulation of natural killer cells in ischemic brain tissues and the chemotactic effect of IP-10. Journal of Neuroinflammation. 2014; 11: 79.
[108]
Zeng X, Zhang YD, Ma RY, Chen YJ, Xiang XM, Hou DY, et al. Activated Drp1 regulates p62-mediated autophagic flux and aggravates inflammation in cerebral ischemia-reperfusion via the ROS-RIP1/RIP3-exosome axis. Military Medical Research. 2022; 9: 25.
[109]
Forrester SJ, Kikuchi DS, Hernandes MS, Xu Q, Griendling KK. Reactive Oxygen Species in Metabolic and Inflammatory Signaling. Circulation Research. 2018; 122: 877–902.
[110]
Bolaños JP, Almeida A. Roles of nitric oxide in brain hypoxia-ischemia. Biochimica et Biophysica Acta. 1999; 1411: 415–436.
[111]
Terpolilli NA, Moskowitz MA, Plesnila N. Nitric oxide: considerations for the treatment of ischemic stroke. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 2012; 32: 1332–1346.
[112]
Wang Y, Hong F, Yang S. Roles of Nitric Oxide in Brain Ischemia and Reperfusion. International Journal of Molecular Sciences. 2022; 23: 4243.
[113]
del Zoppo G, Ginis I, Hallenbeck JM, Iadecola C, Wang X, Feuerstein GZ. Inflammation and stroke: putative role for cytokines, adhesion molecules and iNOS in brain response to ischemia. Brain Pathology (Zurich, Switzerland). 2000; 10: 95–112.
[114]
Narne P, Pandey V, Phanithi PB. Role of Nitric Oxide and Hydrogen Sulfide in Ischemic Stroke and the Emergent Epigenetic Underpinnings. Molecular Neurobiology. 2019; 56: 1749–1769.
[115]
Lakhan SE, Kirchgessner A, Tepper D, Leonard A. Matrix metalloproteinases and blood-brain barrier disruption in acute ischemic stroke. Frontiers in Neurology. 2013; 4: 32.
[116]
Turner RJ, Sharp FR. Implications of MMP9 for Blood Brain Barrier Disruption and Hemorrhagic Transformation Following Ischemic Stroke. Frontiers in Cellular Neuroscience. 2016; 10: 56.
[117]
Hu X, De Silva TM, Chen J, Faraci FM. Cerebral Vascular Disease and Neurovascular Injury in Ischemic Stroke. Circulation Research. 2017; 120: 449–471.
[118]
Sun L, Guo D, Jia Y, Shi M, Yang P, Wang Y, et al. Intercellular adhesion molecule 4 and ischemic stroke: a two-sample Mendelian randomization study. Thrombosis Journal. 2023; 21: 40.
[119]
Koch AE, Halloran MM, Haskell CJ, Shah MR, Polverini PJ. Angiogenesis mediated by soluble forms of E-selectin and vascular cell adhesion molecule-1. Nature. 1995; 376: 517–519.
[120]
El Husseini N, Bushnell C, Brown CM, Attix D, Rost NS, Samsa GP, et al. Vascular Cellular Adhesion Molecule-1 (VCAM-1) and Memory Impairment in African-Americans after Small Vessel-Type Stroke. Journal of Stroke and Cerebrovascular Diseases: the Official Journal of National Stroke Association. 2020; 29: 104646.
[121]
Okada Y, Copeland BR, Mori E, Tung MM, Thomas WS, del Zoppo GJ. P-selectin and intercellular adhesion molecule-1 expression after focal brain ischemia and reperfusion. Stroke. 1994; 25: 202–211.
[122]
Kozuka K, Kohriyama T, Nomura E, Ikeda J, Kajikawa H, Nakamura S. Endothelial markers and adhesion molecules in acute ischemic stroke–sequential change and differences in stroke subtype. Atherosclerosis. 2002; 161: 161–168.
[123]
Bleecker JD, Coulier I, Fleurinck C, Reuck JD. Circulating intercellular adhesion molecule-1 and E-selectin in acute ischemic stroke. Journal of Stroke and Cerebrovascular Diseases: the Official Journal of National Stroke Association. 1998; 7: 192–195.
[124]
Roy S, Das S, Danaboina R, Sharma V, Kaul S, Jyothy A, et al. Association of E-selectin gene polymorphism (S128R) with ischemic stroke and stroke subtypes. Inflammation. 2014; 37: 599–603.
[125]
Vila N, Castillo J, Dávalos A, Esteve A, Planas AM, Chamorro A. Levels of anti-inflammatory cytokines and neurological worsening in acute ischemic stroke. Stroke. 2003; 34: 671–675.
[126]
Spera PA, Ellison JA, Feuerstein GZ, Barone FC. IL-10 reduces rat brain injury following focal stroke. Neuroscience Letters. 1998; 251: 189–192.
[127]
Yang J, Ma K, Zhang C, Liu Y, Liang F, Hu W, et al. Burns Impair Blood-Brain Barrier and Mesenchymal Stem Cells Can Reverse the Process in Mice. Frontiers in Immunology. 2020; 11: 578879.
[128]
Bernardes-Silva M, Anthony DC, Issekutz AC, Perry VH. Recruitment of neutrophils across the blood-brain barrier: the role of E- and P-selectins. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 2001; 21: 1115–1124.
[129]
Konsman JP, Vigues S, Mackerlova L, Bristow A, Blomqvist A. Rat brain vascular distribution of interleukin-1 type-1 receptor immunoreactivity: relationship to patterns of inducible cyclooxygenase expression by peripheral inflammatory stimuli. The Journal of Comparative Neurology. 2004; 472: 113–129.
[130]
Mazzotta G, Sarchielli P, Caso V, Paciaroni M, Floridi A, Floridi A, et al. Different cytokine levels in thrombolysis patients as predictors for clinical outcome. European Journal of Neurology. 2004; 11: 377–381.
[131]
Bö L, Peterson JW, Mørk S, Hoffman PA, Gallatin WM, Ransohoff RM, et al. Distribution of immunoglobulin superfamily members ICAM-1, -2, -3, and the beta 2 integrin LFA-1 in multiple sclerosis lesions. Journal of Neuropathology and Experimental Neurology. 1996; 55: 1060–1072.
[132]
Huang FP, Wang ZQ, Wu DC, Schielke GP, Sun Y, Yang GY. Early NFkappaB activation is inhibited during focal cerebral ischemia in interleukin-1beta-converting enzyme deficient mice. Journal of Neuroscience Research. 2003; 73: 698–707.
[133]
Ohtaki H, Takaki A, Yin L, Dohi K, Nakamachi T, Matsunaga M, et al. Suppression of oxidative stress after transient focal ischemia in interleukin-1 knock out mice. Acta Neurochirurgica. Supplement. 2003; 86: 191–194.
[134]
Wang X, Yue TL, Young PR, Barone FC, Feuerstein GZ. Expression of interleukin-6, c-fos, and zif268 mRNAs in rat ischemic cortex. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 1995; 15: 166–171.
[135]
Azzimondi G, Bassein L, Nonino F, Fiorani L, Vignatelli L, Re G, et al. Fever in acute stroke worsens prognosis. A prospective study. Stroke. 1995; 26: 2040–2043.
[136]
Erta M, Quintana A, Hidalgo J. Interleukin-6, a major cytokine in the central nervous system. International Journal of Biological Sciences. 2012; 8: 1254–1266.
[137]
Hakkoum D, Stoppini L, Muller D. Interleukin-6 promotes sprouting and functional recovery in lesioned organotypic hippocampal slice cultures. Journal of Neurochemistry. 2007; 100: 747–757.
[138]
Tancredi V, D’Antuono M, Cafè C, Giovedì S, Buè MC, D’Arcangelo G, et al. The inhibitory effects of interleukin-6 on synaptic plasticity in the rat hippocampus are associated with an inhibition of mitogen-activated protein kinase ERK. Journal of Neurochemistry. 2000; 75: 634–643.
[139]
Relton JK, Martin D, Thompson RC, Russell DA. Peripheral administration of Interleukin-1 Receptor antagonist inhibits brain damage after focal cerebral ischemia in the rat. Experimental Neurology. 1996; 138: 206–213.
[140]
Idriss HT, Naismith JH. TNF alpha and the TNF receptor superfamily: structure-function relationship(s). Microscopy Research and Technique. 2000; 50: 184–195.
[141]
Zaremba J, Skrobanski P, Losy J. Tumour necrosis factor-alpha is increased in the cerebrospinal fluid and serum of ischaemic stroke patients and correlates with the volume of evolving brain infarct. Biomedicine & Pharmacotherapy. 2001; 55: 258–263.
[142]
Shi X, Yu W, Liu L, Liu W, Zhang X, Yang T, et al. Panax notoginseng saponins administration modulates pro- /anti-inflammatory factor expression and improves neurologic outcome following permanent MCAO in rats. Metabolic Brain Disease. 2017; 32: 221–233.
[143]
Dawson DA, Martin D, Hallenbeck JM. Inhibition of tumor necrosis factor-alpha reduces focal cerebral ischemic injury in the spontaneously hypertensive rat. Neuroscience Letters. 1996; 218: 41–44.
[144]
Lavine SD, Hofman FM, Zlokovic BV. Circulating antibody against tumor necrosis factor-alpha protects rat brain from reperfusion injury. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 1998; 18: 52–58.
[145]
Nawashiro H, Martin D, Hallenbeck JM. Inhibition of tumor necrosis factor and amelioration of brain infarction in mice. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 1997; 17: 229–232.
[146]
Nawashiro H, Martin D, Hallenbeck JM. Neuroprotective effects of TNF binding protein in focal cerebral ischemia. Brain Research. 1997; 778: 265–271.
[147]
Ooboshi H, Ibayashi S, Shichita T, Kumai Y, Takada J, Ago T, et al. Postischemic gene transfer of interleukin-10 protects against both focal and global brain ischemia. Circulation. 2005; 111: 913–919.
[148]
Pelidou SH, Kostulas N, Matusevicius D, Kivisäkk P, Kostulas V, Link H. High levels of IL-10 secreting cells are present in blood in cerebrovascular diseases. European Journal of Neurology. 1999; 6: 437–442.
[149]
Iadecola C, Anrather J. Stroke research at a crossroad: asking the brain for directions. Nature Neuroscience. 2011; 14: 1363–1368.
[150]
McNeill H, Williams C, Guan J, Dragunow M, Lawlor P, Sirimanne E, et al. Neuronal rescue with transforming growth factor-beta 1 after hypoxic-ischaemic brain injury. Neuroreport. 1994; 5: 901–904.
[151]
Zhang L, Wei W, Ai X, Kilic E, Hermann DM, Venkataramani V, et al. Extracellular vesicles from hypoxia-preconditioned microglia promote angiogenesis and repress apoptosis in stroke mice via the TGF-β/Smad2/3 pathway. Cell Death & Disease. 2021; 12: 1068.
[152]
Jickling GC, Liu D, Stamova B, Ander BP, Zhan X, Lu A, et al. Hemorrhagic transformation after ischemic stroke in animals and humans. Journal of Cerebral Blood Flow and Metabolism: Official Journal of the International Society of Cerebral Blood Flow and Metabolism. 2014; 34: 185–199.
[153]
Nguyen QL, Okuno N, Hamashima T, Dang ST, Fujikawa M, Ishii Y, et al. Vascular PDGFR-alpha protects against BBB dysfunction after stroke in mice. Angiogenesis. 2021; 24: 35–46.
[154]
Nurmi A, Lindsberg PJ, Koistinaho M, Zhang W, Juettler E, Karjalainen-Lindsberg ML, et al. Nuclear factor-kappaB contributes to infarction after permanent focal ischemia. Stroke. 2004; 35: 987–991.
[155]
Hayden MS, Ghosh S. Shared principles in NF-kappaB signaling. Cell. 2008; 132: 344–362.
[156]
Ridder DA, Schwaninger M. NF-kappaB signaling in cerebral ischemia. Neuroscience. 2009; 158: 995–1006.
[157]
Luo L, Liu M, Fan Y, Zhang J, Liu L, Li Y, et al. Intermittent theta-burst stimulation improves motor function by inhibiting neuronal pyroptosis and regulating microglial polarization via TLR4/NFκB/NLRP3 signaling pathway in cerebral ischemic mice. Journal of Neuroinflammation. 2022; 19: 141.
[158]
Han HS, Yenari MA. Cellular targets of brain inflammation in stroke. Current Opinion in Investigational Drugs (London, England: 2000). 2003; 4: 522–529.
[159]
Baeuerle PA, Henkel T. Function and activation of NF-kappa B in the immune system. Annual Review of Immunology. 1994; 12: 141–179.
[160]
Russo MA, Sansone L, Carnevale I, Limana F, Runci A, Polletta L, et al. One Special Question to Start with: Can HIF/NFkB be a Target in Inflammation? Endocrine, Metabolic & Immune Disorders Drug Targets. 2015; 15: 171–185.
[161]
Vallabhapurapu S, Karin M. Regulation and function of NF-kappaB transcription factors in the immune system. Annual Review of Immunology. 2009; 27: 693–733.
[162]
Rosenberg GA, Estrada EY, Dencoff JE. Matrix metalloproteinases and TIMPs are associated with blood-brain barrier opening after reperfusion in rat brain. Stroke. 1998; 29: 2189–2195.
[163]
Montaner J, Alvarez-Sabín J, Molina C, Anglés A, Abilleira S, Arenillas J, et al. Matrix metalloproteinase expression after human cardioembolic stroke: temporal profile and relation to neurological impairment. Stroke. 2001; 32: 1789–1766.
[164]
Fujimura M, Gasche Y, Morita-Fujimura Y, Massengale J, Kawase M, Chan PH. Early appearance of activated matrix metalloproteinase-9 and blood-brain barrier disruption in mice after focal cerebral ischemia and reperfusion. Brain Research. 1999; 842: 92–100.
[165]
Yang Y, Rosenberg GA. Matrix metalloproteinases as therapeutic targets for stroke. Brain Research. 2015; 1623: 30–38.
[166]
Hu Q, Chen C, Khatibi NH, Li L, Yang L, Wang K, et al. Lentivirus-mediated transfer of MMP-9 shRNA provides neuroprotection following focal ischemic brain injury in rats. Brain Research. 2011; 1367: 347–359.
[167]
Suzuki Y, Nagai N, Umemura K, Collen D, Lijnen HR. Stromelysin-1 (MMP-3) is critical for intracranial bleeding after t-PA treatment of stroke in mice. Journal of Thrombosis and Haemostasis: JTH. 2007; 5: 1732–1739.
[168]
Kristensen LS, Andersen MS, Stagsted LVW, Ebbesen KK, Hansen TB, Kjems J. The biogenesis, biology and characterization of circular RNAs. Nature Reviews. Genetics. 2019; 20: 675–691.
[169]
Sanger HL, Klotz G, Riesner D, Gross HJ, Kleinschmidt AK. Viroids are single-stranded covalently closed circular RNA molecules existing as highly base-paired rod-like structures. Proceedings of the National Academy of Sciences of the United States of America. 1976; 73: 3852–3856.
[170]
Hsu MT, Coca-Prados M. Electron microscopic evidence for the circular form of RNA in the cytoplasm of eukaryotic cells. Nature. 1979; 280: 339–340.
[171]
Capel B, Swain A, Nicolis S, Hacker A, Walter M, Koopman P, et al. Circular transcripts of the testis-determining gene Sry in adult mouse testis. Cell. 1993; 73: 1019–1030.
[172]
Salzman J, Gawad C, Wang PL, Lacayo N, Brown PO. Circular RNAs are the predominant transcript isoform from hundreds of human genes in diverse cell types. PLoS ONE. 2012; 7: e30733.
[173]
Jeck WR, Sorrentino JA, Wang K, Slevin MK, Burd CE, Liu J, et al. Circular RNAs are abundant, conserved, and associated with ALU repeats. RNA (New York, N.Y.). 2013; 19: 141–157.
[174]
Salzman J, Chen RE, Olsen MN, Wang PL, Brown PO. Cell-type specific features of circular RNA expression. PLoS Genetics. 2013; 9: e1003777.
[175]
Wang PL, Bao Y, Yee MC, Barrett SP, Hogan GJ, Olsen MN, et al. Circular RNA is expressed across the eukaryotic tree of life. PLoS ONE. 2014; 9: e90859.
[176]
Cocquerelle C, Mascrez B, Hétuin D, Bailleul B. Mis-splicing yields circular RNA molecules. FASEB Journal: Official Publication of the Federation of American Societies for Experimental Biology. 1993; 7: 155–160.
[177]
Hansen TB, Jensen TI, Clausen BH, Bramsen JB, Finsen B, Damgaard CK, et al. Natural RNA circles function as efficient microRNA sponges. Nature. 2013; 495: 384–388.
[178]
García-Muse T, Aguilera A. R Loops: From Physiological to Pathological Roles. Cell. 2019; 179: 604–618.
[179]
Liu B, Ye B, Zhu X, Yang L, Li H, Liu N, et al. An inducible circular RNA circKcnt2 inhibits ILC3 activation to facilitate colitis resolution. Nature Communications. 2020; 11: 4076.
[180]
Ashwal-Fluss R, Meyer M, Pamudurti NR, Ivanov A, Bartok O, Hanan M, et al. circRNA biogenesis competes with pre-mRNA splicing. Molecular Cell. 2014; 56: 55–66.
[181]
Conn VM, Hugouvieux V, Nayak A, Conos SA, Capovilla G, Cildir G, et al. A circRNA from SEPALLATA3 regulates splicing of its cognate mRNA through R-loop formation. Nature Plants. 2017; 3: 17053.
[182]
Xia P, Wang S, Ye B, Du Y, Li C, Xiong Z, et al. A Circular RNA Protects Dormant Hematopoietic Stem Cells from DNA Sensor cGAS-Mediated Exhaustion. Immunity. 2018; 48: 688–701.e7.
[183]
Liu CX, Li X, Nan F, Jiang S, Gao X, Guo SK, et al. Structure and Degradation of Circular RNAs Regulate PKR Activation in Innate Immunity. Cell. 2019; 177: 865–880.e21.
[184]
Abdelmohsen K, Panda AC, Munk R, Grammatikakis I, Dudekula DB, De S, et al. Identification of HuR target circular RNAs uncovers suppression of PABPN1 translation by CircPABPN1. RNA Biology. 2017; 14: 361–369.
[185]
Zhao Q, Liu J, Deng H, Ma R, Liao JY, Liang H, et al. Targeting Mitochondria-Located circRNA SCAR Alleviates NASH via Reducing mROS Output. Cell. 2020; 183: 76–93.e22.
[186]
Chen S, Cao X, Zhang J, Wu W, Zhang B, Zhao F. circVAMP3 Drives CAPRIN1 Phase Separation and Inhibits Hepatocellular Carcinoma by Suppressing c-Myc Translation. Advanced Science (Weinheim, Baden-Wurttemberg, Germany). 2022; 9: e2103817.
[187]
Zang J, Lu D, Xu A. The interaction of circRNAs and RNA binding proteins: An important part of circRNA maintenance and function. Journal of Neuroscience Research. 2020; 98: 87–97.
[188]
Rossi F, Beltran M, Damizia M, Grelloni C, Colantoni A, Setti A, et al. Circular RNA ZNF609/CKAP5 mRNA interaction regulates microtubule dynamics and tumorigenicity. Molecular Cell. 2022; 82: 75–89.e9.
[189]
Hafez AK, Zimmerman AJ, Papageorgiou G, Chandrasekaran J, Amoah SK, Lin R, et al. A bidirectional competitive interaction between circHomer1 and Homer1b within the orbitofrontal cortex regulates reversal learning. Cell Reports. 2022; 38: 110282.
[190]
Legnini I, Di Timoteo G, Rossi F, Morlando M, Briganti F, Sthandier O, et al. Circ-ZNF609 Is a Circular RNA that Can Be Translated and Functions in Myogenesis. Molecular Cell. 2017; 66: 22–37.e9.
[191]
Yang Y, Fan X, Mao M, Song X, Wu P, Zhang Y, et al. Extensive translation of circular RNAs driven by N6-methyladenosine. Cell Research. 2017; 27: 626–641.
[192]
Reichert R, Pérez JA, Dalla-Corte A, Pinto E Vairo F, de Souza CFM, Giugliani R, et al. Magnetic resonance imaging findings of the posterior fossa in 47 patients with mucopolysaccharidoses: A cross-sectional analysis. JIMD Reports. 2021; 60: 32–41.
[193]
Zhang Y, Zhang XO, Chen T, Xiang JF, Yin QF, Xing YH, et al. Circular intronic long noncoding RNAs. Molecular Cell. 2013; 51: 792–806.
[194]
Li Z, Huang C, Bao C, Chen L, Lin M, Wang X, et al. Exon-intron circular RNAs regulate transcription in the nucleus. Nature Structural & Molecular Biology. 2015; 22: 256–264.
[195]
Lu Z, Filonov GS, Noto JJ, Schmidt CA, Hatkevich TL, Wen Y, et al. Metazoan tRNA introns generate stable circular RNAs in vivo. RNA (New York, N.Y.). 2015; 21: 1554–1565.
[196]
Mehta SL, Pandi G, Vemuganti R. Circular RNA Expression Profiles Alter Significantly in Mouse Brain After Transient Focal Ischemia. Stroke. 2017; 48: 2541–2548.
[197]
Duan X, Li L, Gan J, Peng C, Wang X, Chen W, et al. Identification and functional analysis of circular RNAs induced in rats by middle cerebral artery occlusion. Gene. 2019; 701: 139–145.
[198]
Liu C, Zhang C, Yang J, Geng X, Du H, Ji X, et al. Screening circular RNA expression patterns following focal cerebral ischemia in mice. Oncotarget. 2017; 8: 86535–86547.
[199]
Li F, Li C, Li X, Li Y, Zhong Y, Ling L. Altered circular RNA expression profiles in the non-ischemic thalamus in focal cortical infarction mice. Aging. 2020; 12: 13206–13219.
[200]
Lu D, Ho ES, Mai H, Zang J, Liu Y, Li Y, et al. Identification of Blood Circular RNAs as Potential Biomarkers for Acute Ischemic Stroke. Frontiers in Neuroscience. 2020; 14: 81.
[201]
Dong Z, Deng L, Peng Q, Pan J, Wang Y. CircRNA expression profiles and function prediction in peripheral blood mononuclear cells of patients with acute ischemic stroke. Journal of Cellular Physiology. 2020; 235: 2609–2618.
[202]
Bai X, Liu X, Wu H, Feng J, Chen H, Zhou D. CircFUNDC1 knockdown alleviates oxygen-glucose deprivation-induced human brain microvascular endothelial cell injuries by inhibiting PTEN via miR-375. Neuroscience Letters. 2022; 770: 136381.
[203]
Jiang Z, Jiang Y. Circular RNA CircPDS5B impairs angiogenesis following ischemic stroke through its interaction with hnRNPL to inactivate VEGF-A. Neurobiology of Disease. 2023; 181: 106080.
[204]
Bai Y, Zhang Y, Han B, Yang L, Chen X, Huang R, et al. Circular RNA DLGAP4 Ameliorates Ischemic Stroke Outcomes by Targeting miR-143 to Regulate Endothelial-Mesenchymal Transition Associated with Blood-Brain Barrier Integrity. The Journal of Neuroscience: the Official Journal of the Society for Neuroscience. 2018; 38: 32–50.
[205]
Zuo L, Xie J, Liu Y, Leng S, Zhang Z, Yan F. Down-regulation of circular RNA CDC14A peripherally ameliorates brain injury in acute phase of ischemic stroke. Journal of Neuroinflammation. 2021; 18: 283.
[206]
Huo H, Hu C, Lu Y, Zhou J, Mai Z. Silencing of circCDC14A prevents cerebral ischemia-reperfusion injury via miR-23a-3p/CXCL12 axis. Journal of Biochemical and Molecular Toxicology. 2022; 36: e22982.
[207]
Li W, Teng J. circCELF1 Induces the Apoptosis and Autophagy of Astrocytes in Ischemic Stroke via Upregulating NFAT5. Cerebrovascular Diseases (Basel, Switzerland). 2023; 52: 306–317.
[208]
Dai Q, Ma Y, Xu Z, Zhang L, Yang H, Liu Q, et al. Downregulation of circular RNA HECTD1 induces neuroprotection against ischemic stroke through the microRNA-133b/TRAF3 pathway. Life Sciences. 2021; 264: 118626.
[209]
Hu T, Li D, Fan T, Zhao X, Chen Z. Circular RNA PUM1 performs as a competing endogenous RNA of microRNA-340-5p to mediate DEAD-box helicase 5 to mitigate cerebral ischemia-reperfusion injury. Bioengineered. 2022; 13: 11564–11578.
[210]
Wang H, Li Z, Gao J, Liao Q. Circular RNA circPTK2 regulates oxygen-glucose deprivation-activated microglia-induced hippocampal neuronal apoptosis via miR-29b-SOCS-1-JAK2/STAT3-IL-1β signaling. International Journal of Biological Macromolecules. 2049; 129: 488–496.
[211]
Huang R, Zhang W, Li W, Gao Y, Zheng D, Bi G. Overexpressing circ_0000831 is sufficient to inhibit neuroinflammation and vertigo in cerebral ischemia through a miR-16-5p-dependent mechanism. Experimental Neurology. 2022; 353: 114047.
[212]
Wang F, Liu J, Wang D, Yao Y, Jiao X. Knockdown of circ_0007290 alleviates oxygen-glucose deprivation-induced neuronal injury by regulating miR-496/PDCD4 axis. Metabolic Brain Disease. 2022; 37: 807–818.
[213]
Ren X, Jing YX, Zhou ZW, Yang JW. Knockdown of circRNA-Memo1 Reduces Hypoxia/Reoxygenation Injury in Human Brain Endothelial Cells Through miRNA-17-5p/SOS1 Axis. Molecular Neurobiology. 2022; 59: 2085–2097.
[214]
Li W, Pang H, Xie L. Depletion of circ_0006459 protects human brain microvascular endothelial cells from oxygen-glucose deprivation-induced damage through the miR-940/FOXJ2 pathway. Transplant Immunology. 2023; 80: 101780.
[215]
Liu D, Xiao H, Liu J, Zhang Y, Li J, Zhang T, et al. Circ_0000566 contributes oxygen-glucose deprivation and reoxygenation (OGD/R)-induced human brain microvascular endothelial cell injury via regulating miR-18a-5p/ACVR2B axis. Metabolic Brain Disease. 2023; 38: 1273–1284.
[216]
Dai Y, Sheng Y, Deng Y, Wang H, Zhao Z, Yu X, et al. Circ_0000647 promotes cell injury by modulating miR-126-5p/TRAF3 axis in oxygen-glucose deprivation and reperfusion-induced SK-N-SH cell model. International Immunopharmacology. 2022; 104: 108464.
[217]
Pei L, Xu X, Yuan T. Circ_0101874 overexpression strengthens PDE4D expression by targeting miR-335-5p to promote neuronal injury in ischemic stroke. Journal of Stroke and Cerebrovascular Diseases: the Official Journal of National Stroke Association. 2022; 31: 106817.
[218]
Qiu L, He J, Chen H, Xu X, Tao Y. CircDLGAP4 overexpression relieves oxygen-glucose deprivation-induced neuronal injury by elevating NEGR1 through sponging miR-503-3p. Journal of Molecular Histology. 2022; 53: 321–332.
[219]
Cui Y, Wang Y, Song X, Ning H, Zhang Y, Teng Y, et al. Brain endothelial PTEN/AKT/NEDD4-2/MFSD2A axis regulates blood-brain barrier permeability. Cell Reports. 2021; 36: 109327.
[220]
Zuo L, Zhang L, Zu J, Wang Z, Han B, Chen B, et al. Circulating Circular RNAs as Biomarkers for the Diagnosis and Prediction of Outcomes in Acute Ischemic Stroke. Stroke. 2020; 51: 319–323.
[221]
Xin H, Katakowski M, Wang F, Qian JY, Liu XS, Ali MM, et al. MicroRNA cluster miR-17-92 Cluster in Exosomes Enhance Neuroplasticity and Functional Recovery After Stroke in Rats. Stroke. 2017; 48: 747–753.
[222]
Zhu X, Ding J, Wang B, Wang J, Xu M. Circular RNA DLGAP4 is down-regulated and negatively correlates with severity, inflammatory cytokine expression and pro-inflammatory gene miR-143 expression in acute ischemic stroke patients. International Journal of Clinical and Experimental Pathology. 2019; 12: 941–948.
[223]
Huang Z, Xu Y, Wan M, Zeng X, Wu J. miR-340: A multifunctional role in human malignant diseases. International Journal of Biological Sciences. 2021; 17: 236–246.
[224]
Li Z, Wong KY, Calin GA, Chng WJ, Chan GCF, Chim CS. Epigenetic silencing of miR-340-5p in multiple myeloma: mechanisms and prognostic impact. Clinical Epigenetics. 2019; 11: 71.
[225]
Novosadova E, Chabronova A, Kolek V, Petrek M, Navratilova Z. The Serum Expression of Selected miRNAs in Pulmonary Sarcoidosis with/without Löfgren’s Syndrome. Mediators of Inflammation. 2016; 2016: 1246129.
[226]
Bradová M, Tejnecký V, Borůvka L, Němeček K, Ash C, Šebek O, et al. The variations of aluminium species in mountainous forest soils and its implications to soil acidification. Environmental Science and Pollution Research International. 2015; 22: 16676–16687.
[227]
Ren K, Yu Y, Wang X, Liu H, Zhao J. MiR-340-3p-HUS1 axis suppresses proliferation and migration in lung adenocarcinoma cells. Life Sciences. 2021; 274: 119330.
[228]
Yang L, Men WL, Yan KM, Tie J, Nie YZ, Xiao HJ. MiR-340-5p is a potential prognostic indicator of colorectal cancer and modulates ANXA3. European Review for Medical and Pharmacological Sciences. 2018; 22: 4837–4845.
[229]
Ma Y, Shan Z, Ma J, Wang Q, Chu J, Xu P, et al. Validation of downregulated microRNAs during osteoclast formation and osteoporosis progression. Molecular Medicine Reports. 2016; 13: 2273–2280.
[230]
Waiboci LW, Ahmed CM, Mujtaba MG, Flowers LO, Martin JP, Haider MI, et al. Both the suppressor of cytokine signaling 1 (SOCS-1) kinase inhibitory region and SOCS-1 mimetic bind to JAK2 autophosphorylation site: implications for the development of a SOCS-1 antagonist. Journal of Immunology (Baltimore, Md.: 1950). 2007; 178: 5058–5068.
[231]
Baek SH, Ko JH, Lee H, Jung J, Kong M, Lee JW, et al. Resveratrol inhibits STAT3 signaling pathway through the induction of SOCS-1: Role in apoptosis induction and radiosensitization in head and neck tumor cells. Phytomedicine: International Journal of Phytotherapy and Phytopharmacology. 2016; 23: 566–577.
[232]
Samavati L, Rastogi R, Du W, Hüttemann M, Fite A, Franchi L. STAT3 tyrosine phosphorylation is critical for interleukin 1 beta and interleukin-6 production in response to lipopolysaccharide and live bacteria. Molecular Immunology. 2009; 46: 1867–1877.
[233]
Tan PH, Tyrrell HEJ, Gao L, Xu D, Quan J, Gill D, et al. Adiponectin receptor signaling on dendritic cells blunts antitumor immunity. Cancer Research. 2014; 74: 5711–5722.
[234]
Li Q, Tian Z, Wang M, Kou J, Wang C, Rong X, et al. Luteoloside attenuates neuroinflammation in focal cerebral ischemia in rats via regulation of the PPARγ/Nrf2/NF-κB signaling pathway. International Immunopharmacology. 2019; 66: 309–316.
[235]
Rizzo MR, Fasano R, Paolisso G. Adiponectin and Cognitive Decline. International Journal of Molecular Sciences. 2020; 21: 2010.
[236]
Duan S, Wang F, Cao J, Wang C. Exosomes Derived from MicroRNA-146a-5p-Enriched Bone Marrow Mesenchymal Stem Cells Alleviate Intracerebral Hemorrhage by Inhibiting Neuronal Apoptosis and Microglial M1 Polarization. Drug Design, Development and Therapy. 2020; 14: 3143–3158.
[237]
Chen C, Chang X, Zhang S, Zhao Q, Lei C. CircRNA CTNNB1 (circCTNNB1) ameliorates cerebral ischemia/reperfusion injury by sponging miR-96-5p to up-regulate scavenger receptor class B type 1 (SRB1) expression. Bioengineered. 2022; 13: 10258–10273.
[238]
Zhou J, Du T, Li B, Rong Y, Verkhratsky A, Peng L. Crosstalk Between MAPK/ERK and PI3K/AKT Signal Pathways During Brain Ischemia/Reperfusion. ASN Neuro. 2015; 7: 1759091415602463.
[239]
De S, Dermawan JKT, Stark GR. EGF receptor uses SOS1 to drive constitutive activation of NFκB in cancer cells. Proceedings of the National Academy of Sciences of the United States of America. 2014; 111: 11721–11726.
[240]
Sawe N, Steinberg G, Zhao H. Dual roles of the MAPK/ERK1/2 cell signaling pathway after stroke. Journal of Neuroscience Research. 2008; 86: 1659–1669.
[241]
Tan J, Liu D, Lv X, Wang L, Zhao C, Che Y, et al. MAPK mediates inflammatory response and cell death in rat pulmonary microvascular endothelial cells in an ischemia-reperfusion model of lung transplantation. The Journal of Heart and Lung Transplantation: the Official Publication of the International Society for Heart Transplantation. 2013; 32: 823–831.
[242]
Pérez-Sánchez C, Gómez-Ferrería MA, de La Fuente CA, Granadino B, Velasco G, Esteban-Gamboa A, et al. FHX, a novel fork head factor with a dual DNA binding specificity. The Journal of Biological Chemistry. 2000; 275: 12909–12916.
[243]
Tang Y, Zhou M, Huang R, Shen L, Yang L, Zhou Z, et al. Involvement of HECTD1 in LPS-induced astrocyte activation via σ-1R-JNK/p38-FOXJ2 axis. Cell & Bioscience. 2021; 11: 62.
[244]
Xu X, Wu Z, Qiu H, Wu J. Circular RNA circPHC3 Promotes Cell Death and Apoptosis in Human BMECs After Oxygen Glucose Deprivation via miR-455-5p/TRAF3 Axis in vitro. Neuropsychiatric Disease and Treatment. 2021; 17: 147–156.
[245]
Feng B, Meng L, Luan L, Fang Z, Zhao P, Zhao G. Upregulation of Extracellular Vesicles-Encapsulated miR-132 Released From Mesenchymal Stem Cells Attenuates Ischemic Neuronal Injury by Inhibiting Smad2/c-jun Pathway via Acvr2b Suppression. Frontiers in Cell and Developmental Biology. 2021; 8: 568304.
[246]
Chen W, Wang H, Feng J, Chen L. Overexpression of circRNA circUCK2 Attenuates Cell Apoptosis in Cerebral Ischemia-Reperfusion Injury via miR-125b-5p/GDF11 Signaling. Molecular Therapy. Nucleic Acids. 2020; 22: 673–683.
[247]
Ohshiro K, Chen J, Srivastav J, Mishra L, Mishra B. Alterations in TGF-β signaling leads to high HMGA2 levels potentially through modulation of PJA1/SMAD3 in HCC cells. Genes & Cancer. 2020; 11: 43–52.
[248]
Peng L, Yin J, Wang S, Ge M, Han Z, Wang Y, et al. TGF-β2/Smad3 Signaling Pathway Activation Through Enhancing VEGF and CD34 Ameliorates Cerebral Ischemia/Reperfusion Injury After Isoflurane Post-conditioning in Rats. Neurochemical Research. 2019; 44: 2606–2618.
[249]
Zhao Y, Zhang X, Chen X, Wei Y. Neuronal injuries in cerebral infarction and ischemic stroke: From mechanisms to treatment (Review). International Journal of Molecular Medicine. 2022; 49: 15.
[250]
Sakai S, Shichita T. Inflammation and neural repair after ischemic brain injury. Neurochemistry International. 2019; 130: 104316.
[251]
Jiang YJ, Cao SQ, Gao LB, Wang YY, Zhou B, Hu X, et al. Circular Ribonucleic Acid Expression Profile in Mouse Cortex after Traumatic Brain Injury. Journal of Neurotrauma. 2019; 36: 1018–1028.
[252]
Siracusa C, Sabatino J, Leo I, Eyileten C, Postuła M, De Rosa S. Circular RNAs in Ischemic Stroke: Biological Role and Experimental Models. Biomolecules. 2023; 13: 214.
[253]
Gasull T, Arboix A. Molecular Mechanisms and Pathophysiology of Acute Stroke: Emphasis on Biomarkers in the Different Stroke Subtypes. International Journal of Molecular Sciences. 2022; 23: 9476.

Publisher’s Note: IMR Press stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share
Back to top