IMR Press / FBL / Volume 28 / Issue 9 / DOI: 10.31083/j.fbl2809208
Open Access Review
Role of Ferroptosis in Glial Cells after Ischemic Stroke
Show Less
1 Department of Acupuncture and Rehabilitation, Affiliated Hospital of Nanjing University of Chinese Medicine, 210029 Nanjing, Jiangsu, China
*Correspondence: pengyongjun2004@126.com (Yong-Jun Peng)
Front. Biosci. (Landmark Ed) 2023, 28(9), 208; https://doi.org/10.31083/j.fbl2809208
Submitted: 10 February 2023 | Revised: 6 April 2023 | Accepted: 9 May 2023 | Published: 19 September 2023
Copyright: © 2023 The Author(s). Published by IMR Press.
This is an open access article under the CC BY 4.0 license.
Abstract

Ischemic stroke (IS) is the predominant cause of morbidity and mortality worldwide. Ferroptosis, a new type of programmed cell death, has been shown to play a crucial role in IS pathogenesis. Traditionally, research focused on neurons did not uncover specific positive results for IS. However, glial cells have recently received interest as promising targets for IS treatment, not only for their structural function but also in the iron transfer between glia and neurons, which indicates a promising glia–neuron crosstalk in mediating the IS process and ischemia/reperfusion-associated neuropathology, showing their affiliation with ferroptosis. This review addresses the major phenomena of iron metabolism and the process and regulation of ferroptosis, with a particular focus on their impact on IS pathology. The review discusses iron homeostasis, the biology of reactive oxygen species, and lipid peroxidation for modulating the process of IS-induced ferroptosis in different glial cells. We then review recent therapies that leverage ferroptosis modulation for the treatment of IS. Extensive preclinical and clinical research is necessary to fully understand the roles of glia–neuron crosstalk and ferroptosis in IS.

Keywords
glia
neuron
ferroptosis
ischemic stroke
iron
crosstalk
1. Introduction

Stroke is a predominant cause of disability and death worldwide [1]. Of all stroke types, the most common is ischemic stroke (IS) and evidence based treatments such as intravenous thrombolysis with recombinant tissue plasminogen activator and endovascular thrombectomy are the most effective treatments, currently with a low proportion of eligible patients and a short time window [2]. Hence, more effective targets for IS treatment are required and need further investigation. Potential mechanisms of IS pathogenesis have been gradually uncovered, including neuroinflammation, metal ion imbalance, mitochondrial dysfunction, oxidative injury, and lipid peroxidation [3, 4]. Although research focused on neurons has not presented a high rate of positive results for IS, glia cells have recently received interest as promising therapeutic effectors for IS treatments. Activated glia have multiple functions because they can release neurotrophic factors and proinflammatory cytokines and phagocytose with different pathways during the onset and progression of IS [5]. Clinical studies have reported that high serum ferritin levels result in a poor IS prognosis [6, 7, 8, 9]. Notably, ferroptosis is an iron-dependent programmed cell death (PCD) and have been demonstrated to play a role in regulating iron metabolism and lipid peroxidation. Recently, researchers have proven that suppressing ferroptosis reverses ischemic injuries [10, 11, 12]. In this review, we discuss iron homeostasis, the biology of reactive oxygen species (ROS), and lipid peroxidation for modulating IS-induced ferroptosis. We also address iron metabolism, the mechanism of ferroptosis in glial cells and neurons, and their impact on IS pathologies. Lastly, we review recent therapies that focus on ferroptosis for IS treatment.

2. Evidence of Ferroptosis in IS
2.1 Ferroptosis Mechanisms

Ferroptosis is a regulated and iron-dependent form of PCD that is distinct from apoptosis, necroptosis, autophagy, and other forms of cell death [13]. With respect to its morphology, ferroptosis targets mitochondria, which present a small volume, increases membrane density, reduces or vanishes crista, and thus ruptures the outer membrane while the nuclei structure stays intact [13, 14, 15]. By contrast, margination, condensation, and fragmentation of chromatin, as well as the generation of apoptotic bodies and plasma membrane blebbing, are typically involved in apoptosis [16]. The characteristics of necroptosis are membrane rupture and the production of membrane bubbles, which are stained positive for annexin V [17]. As another form of PCD, autophagy is characterized by the de novo synthesis of cellular components such as macroproteins or organelles sequestered into a double membrane and fused with a lysosome for digestion and recycling [18]. These characteristics are not present in ferroptosis, which has unique morphological characteristics.

Iron accumulation and excessive lipid peroxidation are the primary bioenergetic features of ferroptosis (Fig. 1). It has been demonstrated that three major mechanisms may control ferroptosis: iron regulation, ROS biology, and metabolism [19].

Fig. 1.

Possible signaling pathway of ferroptosis. Ferroptosis occurs in two major pathways: (1) through the accumulation of ferrous ions via iron-transport molecules, such as transferrin, transferrin receptor, and ferroprotein; and (2) through the accumulation of lipid peroxides. ACSL4, acyl-CoA synthetase long-chain family member 4; CoQ10, coenzyme Q10; CP, ceruloplasmin; DMT1, divalent metal transporter 1; ETC, electron transport chain; FPN1, ferroportin 1; GPX4, glutathione peroxidase 4; GR, glutathione reductase; GSH, glutathione; GSSG, glutathione disulfide; HNE, hydroxynonenal; HO-1, heme oxygenase 1; LOX, lipoxygenase; LPCAT3, lysophosphatidylcholine acyltransferase 3; MDA, malondialdehyde; NCOA4, nuclear receptor co-activator 4; PUFA, polyunsaturated fatty acids; Se, selenium; SLC3A2, subunit solute carrier family 3 member 2; SLC7A11, solute carrier family 7 member 11; STEAP3, six-transmembrane epithelial antigen of prostate 3; Tf, transferrin; TfR1, transferrin receptor 1.

Although the exact molecular event of neuronal ferroptosis remains obscure, the contribution of glial cells to mediate iron-induced toxicity to neurons has been uncovered. It is very well known that neuroglia contributes to the maintenance of brain iron metabolism and physiological function. Moreover, glia express and distribute almost all iron transporters and various iron metabolism proteins. Accumulated studies have identified glia as essential for regulating the ferroptosis of neuronal death [20, 21].

2.2 Ferroptosis is a Prominent Pathway of Cell Death in IS

Ferroptosis is involved in the pathological process of IS, which is a novel type of cell death after ischemia onset [22]. Table 1 (Ref. [12, 23, 24, 25, 26, 27, 28, 29, 30, 31, 32, 33, 34, 35]) presents several targets of ferroptosis genes that are associated with IS. In addition, the characteristics of ferroptosis are highly consistent with the pathological changes in patients with IS. For instance, the prognosis of IS is worse in older patients, who have cerebral iron deposition, indicating that iron accumulation may exacerbate ferroptosis of neurons after IS injury [6]. Moreover, an association between high ferritin levels and poorI S prognosis has been found [7]. Gill et al. [8] reported that according to observational studies, a higher transferrin level was associated with a decreased stroke risk. It was also shown that the level of the iron-regulatory hormone hepcidin was significantly higher in patients with IS [9]. Additionally, ferroptosis occurred in different IS animal and cellular models, such as middle cerebral artery occlusion/reperfusion (MCAO/R) rats and PC12 cells treated with oxygen glucose deprivation and reperfusion (OGD/R), including iron accumulation [23, 36], glutathione (GSH) [24, 37, 38], lipid peroxidation [25, 26, 39], oxidative stress [27, 40], and coenzyme Q10 (CoQ10) [41]. Edaravone, as the only drug among ferroptosis inhibitors, has been approved for acute IS therapy [42, 43] and demonstrated to suppress ferroptosis in the ischemic cerebral area. Other drugs were only found to be effective preclinically, such as deferoxamine [44] and selenium [11], but clinical tests are still needed. Although evidence indicates that iron homeostasis and ferroptosis play a prominent role in mediating neuronal death in IS, how these mechanisms occur at each stage of stroke remains largely unknown. Research on the biological processes targeting iron accumulation and associated cell death is urgently needed.

Table 1.Ferroptosis-related genes in ischemic stroke.
Gene Function Model Refs
PVT1 PVT1 mediates ferroptosis via miR-214 and by regulating p53 and TfR1 levels. Patients with AIS and OGD/R PC12 cells [23]
ACSL4 ACSL4 promotes ferroptosis via enhancing lipid peroxidation and contributes to microglia-mediated inflammation. MCAO [25]
Inhibiting ACSL4 upregulates GPX4 expression in ischemic stroke, suppresses iron accumulation, and decreases GPX activity to inhibit ferroptosis. MCAO [12]
NCOA4 The NCOA4 deletion abrogates ischemic stroke-induced ferritinophagy and suppresses ferroptosis both in vivo and in vitro. MCAO and OGD/R [28]
Tau Tau promotes iron export to stabilize Fpn, which results in iron accumulation. MCAO [29]
9a 9a blocks ferroptosis via disrupting the NCOA4-FTH1 protein-protein interaction. MCAO [30]
PGE2 PGE2 inhibits Fe2+, glutathione oxidation, and lipid peroxidation to suppress ferroptosis. Patients with cranial hypertension and MCAO. [31]
LncRNA Meg3 Treatment with Meg3-siRNA inhibits ferroptotic events through regulating GPX4 expression at ischemic onset. OGD/R RBMVECs [24]
cPKCγ The cPKCγ deletion decreases GPX4, COX2, and ACSL4 expressions. OGD/R primary cultured cortical neurons [26]
UBIAD1 UBIAD1 significantly suppresses ferroptosis via elevating SOD, T-AOC, the production of CoQ10, and eNOS-regulated NO generation. MCAO and OGD/R. [27]
PTRF PTRF targets PLA2G4A to promote lipid metabolism and alter mitochondrial bioenergetics and ferroptosis in neurons. MCAO and OGD/R HT22 and BV2 cells [32]
PMU2 PMU2 exacerbates neuroinflammation via SLC7A11-dependent inhibition of ferroptosis at ischemic stroke onset. MCAO/R and cells with OGD/R-induced cortical neuron injury [33]
TLR4 TLR4 induces neuronal ferroptosis through promoting oxidative stress and neuroinflammation. HIBD and OGD/R HT22 cells [34]
circ-Carm1 circ-Carm1 stimulates ferroptosis. Patients with ACI and OGD/R HT22 cells [35]

PVT1, plasmacytoma variant 1; Fpn, ferroportin; GPX4, Glutathione peroxidase 4; NCOA4, nuclear receptor coactivator 4; FTH1, Porcine Ferritin heavy chain; PGE2, prostaglandin E2; RBMVEC, rat brain microvascular endothelial cells; OGD, oxygen-glucose deprivation; cPKCγ, conventional protein kinase cγ; COX2, cyclooxygenase 2; UBIAD1, UbiA prenyltransferase domain containing 1; SOD, superoxide dismutase; CoQ10, coenzyme Q10; eNOS, endothelial nitric oxide synthase; NO, nitric oxide; PTRF, polymerase I and transcript release factor; PMU2, pumilio 2; SIRT1, sirtuin 1; TLR4, toll-like receptor 4; MCAO, middle cerebral artery occlusion and reperfusion; HIBD, hypoxic-ischemic brain damage.

3. Ferroptosis Induced by Glial Cells

Glial activation is involved in the ischemic cascade reaction of IS through various pathways (Fig. 2). Activation of microglia and astrocytes, which are typically classified into M1/A1 and M2/A2 phenotypes, is one of the earliest events after brain ischemia [45, 46, 47]. M1/A1 and M2/A2 release pro- and anti-inflammatory cytokines, respectively, to present neurotoxic and neuroprotective functions after brain ischemia [47, 48]. Oligodendrocytes are known to produce superoxide radicals, lipid peroxidation, and iron oxidation at ischemia onset [49]. Different types of glia may have effects on IS pathogenesis via various pathways, and glial activation may also affect neuronal ferroptosis via iron homeostasis, lipid metabolism, and oxidative damage.

Fig. 2.

Possible neuron–glia crosstalk in ischemia-induced ferroptosis. HIF-1α promotes iron accumulation in neurons and glia by upregulating DMT1 and ferrous uptake in hypoxia conditions. In microglia, HIF-1α inhibits ACSL4 expression, which can promote ROS preposition. Additionally, HIF-1α upregulates inflammatory cytokines, such as IL-6 and TNF-α in astrocytes, which can further upregulate DMT1, downregulate FPN1, and results in iron accumulation in neurons. ROS released from glia result in neuronal oxidative stress. In astrocytes, upregulation of Nrf2 induces GCL expression, increases GSH synthesis, and inhibits ROS deposition, which all alleviate neuronal oxidative stress. ACSL4, acyl-CoA synthetase long-chain family member 4; ATP, adenosine triphosphate; CP, ceruloplasmin; DMT1, divalent metal transporter 1; Fer-1, Ferrostatin-1; FPN1, ferroportin 1; HIF-1α, Hypoxia-Inducible Factor 1-Alpha; Heph, hephaestin; IL-6, interleukin 6; GAST, glutamate-aspartate transporter; GCL, glutamate-cysteine ligase; Gln, glutamine; GLT, glutamate transporter; Glu, glutamate; GPI, glycosylphosphatidylinositol; GSH, glutathione; Nrf2, nuclear factor erythroid 2-related factor 2; NOX, NADPH-oxidase; ROS, reactive oxygen species; SLC3A2, subunit solute carrier family 3 member 2; SLC7A11, solute carrier family 7 member 11; Tf, transferrin; TfR1, transferrin receptor 1; TNF-α, tumor necrosis factor α.

3.1 Iron Regulation Induces Ferroptosis in Glial Cells after IS

It is known that iron can cross the blood-brain barrier (BBB) and be released into the brain’s interstitial fluid. Neurons and glial cells can absorb and unevenly distribute iron [50]. Studies have confirmed that almost all the iron transporters and proteins involved in iron metabolism are expressed in glia [51, 52, 53]. It has been reported that iron retention is much higher in microglia than in neurons [54]. Hence, glial cells are critical to maintaining iron homeostasis and physiological and pathological function in the brain. In addition, ferritin is directly driven by hypoxia [55], indicating that glial cells have effects on IS via regulating iron metabolism.

3.1.1 Microglia

Microglia serve a key role in the regulation of iron homeostasis in the central nervous system (CNS) [56]. Intuitively, iron homeostasis relies on the expression of iron-uptake and iron-release proteins on the cellular membrane. The mechanism of iron uptake in microglia has been uncovered, and microglial iron accumulation occurs in cerebral ischemia. It was found that microglia are able to take up non-transferrin-bound iron (NTBI) via a transferrin-independent mechanism [57, 58]. Moreover, transferrin receptor 1 (TfR1) has been identified in microglia both in vivo and in vitro [59, 60, 61]. It was reported that the expression of TfR1 and intracellular iron are both upregulated in hypoxia-induced microglia [60]. Divalent metal transporter 1 (DMT1), a pivotal protein for both transferrin-dependent and transferrin-independent iron uptake, has been identified in microglia [56, 62, 63, 64]. Hypoxia-inducible factor-1 (HIF-1) is known as a critical regulator, which is responsible for the facilitation, adaptation, and survival of cells from normoxia to hypoxia [65, 66]. Qian et al. [67] demonstrated that DMT1 is a target gene of HIF-1 and that silencing HIF-1 down-regulates DMT1 expression and ferrous uptake. Hence, the identification of TfR1, DMT1, and iron content in hypoxia indicates that microglia might take up transferrin-Fe via NTBI and TfR1/DMT1 pathways and might be regulated in hypoxia. Ferroportin 1 (Fpn1) is currently the only known microglial iron exporter [68]. It was also verified that Hephaestin (Heph), which is a critical ferroxidase, is expressed in microglia [69]. Yang et al. [70] reported that the levels of Fpn1 and Heph proteins were downregulated in MCAO models. Another study found that ischemic onset significantly upregulated the levels of the iron-import proteins TfR1 and DMT1 and decreased the expression of Fpn1 in reactive microglia, also indicating that iron deposition occurred in activated microglia after IS [71]. Thus, Fpn1/Heph may be the pathway of iron efflux in physiological and ischemia-induced microglia. Moreover, accumulated iron in hypoxia-induced microglia can stimulate ROS and modulate inflammatory processes via the generation of pro-inflammatory and anti-inflammatory cytokines [72]. In hypoxic-ischemic conditions, an excessive gliosis process can result in tertiary cerebral injury, and astrocyte cytotoxicity would be induced by microglial signaling, while the regular microglial response promotes the regeneration process. On the other hand, studies suggest that microglia activation prevents excessive iron accumulation in response to hypoxic damage [73, 74]. It was reported that excessively retained microglial iron is a key mediator of oligodendrocyte cytotoxicity in periventricular white matter following hypoxia [74]. Taken together, activated microglia play a crucial role in physiological and cerebral ischemia via regulating iron metabolism, and this has been partially corroborated, implying that a treatment strategy targeting microglial iron metabolism and ferroptosis modulation may have potential therapeutic value in treating IS.

3.1.2 Astrocytes

Astrocytes, one of the most important components of the BBB, influence the regulation of iron homeostasis in the cerebral area [75, 76]. Studies have shown that astrocytes are responsible for acquiring nutrients such as iron from the circulating blood [77]. Astrocytes are also able to transport various types of iron, such as Tf-Fe [78], NTBI [79], and heme [80], through coupling with proteins such as ceruloplasmin (CP), which is known as the form of glycosylphosphatidylinositol-anchored proteins in the astrocyte membrane [81, 82, 83]. In addition, astrocytes can take up iron via the TfR1/DMT1 pathway, and it is transferred to proteins, including Fpn1 and CP [84]. Moreover, findings suggest that iron effluxes from astrocytes via the Fpn1/CP pathway [85]. CP is highly expressed in astrocytes and suppresses iron-induced lipid peroxidation [86]. Patients with aceruloplasminemia present serious iron accumulation in astrocytes and iron deposition, which is also observed in the astrocytes of mice consistently lacking CP [87, 88]. Ryan et al. [89] reported that the CP level was significantly upregulated in permanent MCAO models. Conversely, loss of CP dilated the lesion size and aggravated neurological function recovery with a decline in ferroportin and a raise in DMT1 proteins, indicating the neuroprotective effects of CP in IS via iron regulation in astrocytes [89]. In addition, hypoxia can stimulate the expression of HIF-1α, as well as TfR1, DMT1, and Fpn1 proteins, thus promoting Tf-Fe and NTBI uptake and iron release in astrocytes and a progressive enhancement in cellular iron content [90]. Therefore, strategies targeting iron metabolism in activated astrocytes might play a role in IS pathogenesis.

3.1.3 Oligodendrocytes

Oligodendrocytes, a myelin-forming cell type in the cerebral area, have high iron requirements for normal function. Moreover, they require two- to three-fold higher energy levels than other cell types in the CNS during myelination. Oligodendrocytes also synthesize cholesterol, which is a highly metabolically demanding process and makes them vulnerable to hypoperfusion. The white matter, which is comprised of critical axonal projections, is especially vulnerable to ischemia and has attracted attention lately. About 95% of patients with IS suffer from white matter injury, accounting for half of the stroke total mean infarct volume [91]. It has been considered that white matter injury after stroke is highly correlated with axon degeneration, oligodendrocyte damage, and nerve demyelination [92]. A recent study showed that ischemia increases intracellular iron uptake with the upregulation of oligodendrocytes and myelin markers (markers of oligodendrocyte differentiation [MBP], oligodendrocyte-specific protein [OSP], and helix-loop-helix transcription factor) during the recovery phase of IS [93]. Michalski et al. [94] reported that OSP and MBP display early expression after the ischemic onset and promote protective and regenerative reactions in IS pathogenesis. Therefore, oligodendrocytes might have effects on white matter injury, such as myelination, after IS via iron regulation.

3.2 ROS-Derived Ferroptosis in Glial Cells

It is known that ferroptosis is ultimately driven by oxidative stress [19]. After IS, the damage-associated molecular patterns are activated, inducing a sequence of pathological changes such as oxidative stress [95]. Oxidative stress refers to the relative excess of ROS induced by disrupting the balance between ROS production and depletion. Recent studies have shown that glial cells are involved in ROS regulation, which might further induce ferroptosis in IS [96].

3.2.1 Microglia

Activated microglia are frequently present during oxidative stress and the inflammatory processes of IS [97]. Specifically, they are closely connected to an excessive ROS response in the early stage of IS [98, 99]. NADPH oxidase (NOX), which is known as a primary source of ROS production, has been consistently reported to participate in the pathogenesis of cerebral ischemic injury [100]. Recently, NOX was also identified as a biomarker predicting ferroptosis via GPX4-independent control of the lipid ROS accumulation system [101, 102]. During ischemic conditions, NOX-mediated ROS production is reported to occur in microglia, and targeting NOX-dependent ROS generation suppresses microglial activation and polarization and ferroptosis [103]. Peroxiredoxin-1 (Prdx1) is known as one of the typical antioxidant enzymes, and it can inhibit ROS deposition and keep the reduction-oxidation status in both normal oxidative metabolism and oxidative stress [104, 105]. In a recent study, Kim et al. [106] identified stroke-associated microglia (SAM) and showed that Prdx1 deficiency decreased the extent of SAM in IS and exacerbated cerebral injury, which indicates that SAM might have potential antioxidant effects on stroke-induced microglia. Under neuroinflammatory conditions, reduced ROS levels and fewer apoptotic neurons present in CX3CR1-deficient MCAO mice suppressed the proliferation and regulated the polarization of microglia, which inhibited the process of releasing inflammatory cytokines to improve the ischemic brain injury outcome [107]. Consistently, inhibiting the production of microglial proinflammatory cytokines and lipid peroxidation via silencing acyl-CoA synthetase long-chain family member 4 (ACSL4) in IS has proved to inhibit ferroptosis in neurons [108].

3.2.2 Astrocytes

Inhibiting excessive ROS production after IS has been proposed as the most effective strategy to improve the prognosis outcome. Elastin-derived peptides are shown as the scavengers of free radicals via the induced activities of antioxidant enzymes, which are detectable in the cerebrospinal fluid of patients after IS [109, 110]. In mouse astrocytes induced by OGD, the Val-Gly-Val-Ala-Pro-Gly (VGVAPG) peptide upregulated nitric oxide release during ROS accumulation [111]. However, the effects were reversed by decreasing ROS production when knocking down the galactosidase beta 1 gene, which indicates that the VGVAPG peptide has pro-survival effects on activating astrocytes via mediating ROS regulation [111].

The functional impairment of mitochondria, resulting in increased ROS deposition and oxidative stress in cerebral ischemia, has been shown to initiate and enhance the vulnerability of cells to ferroptosis [19, 112]. Some important modulators, such as triiodo-l-thyronine and inositol trisphosphate receptor type 2, which can promote astrocyte mitochondrial adenosine triphosphate (ATP) production and the disruption of astrocytic Ca2+ signaling, respectively, have been proposed to have effects on mitochondrial dysfunction [113, 114].

3.2.3 Oligodendrocytes

Oligodendrocytes and myelin damage cause the functional impairment in IS. Glutamate, which has been recently proposed to promote ferroptosis [115], also contributes to oligodendrocyte excitotoxicity and axonal dysfunction during IS onset [116, 117]. In addition, oligodendrocytes are known to express purinergic receptors, and ATP release and P2X7 receptor activation result in white matter injury in ischemia [118, 119]. Ca2+ influx is the core event in glutamate and ATP excitotoxicity, which further contributes to increased ROS production [118, 120, 121]. It has been reported that ferroptosis contributes to spinal cord injury-induced white matter damage, and ferrostatin-1, a potent inhibitor of ferroptosis, was demonstrated to inhibit excessive ROS deposition and reduce the expression of ferroptosis-related genes and their products. The products of these genes are iron-responsive element binding protein 2 and prostaglandin endoperoxide synthase 2, which can further suppress ferroptosis in oligodendrocytes, thus alleviating white matter injury and promoting functional reinstitution. Plenty of research indicates that oligodendrocytes might be involved in IS via ROS-induced ferroptosis, but further investigations are needed.

3.3 Lipid Metabolism-Derived Ferroptosis in Glial Cells

After IS, iron accumulation and ROS deposition can further contribute to lipid peroxidation via the Fenton reaction, which has been proposed as the vital lethal factor inducing ferroptosis [122]. Studies focusing on the suppression of lipid peroxidation-induced ferroptosis might provide a new target for IS therapies [29, 123, 124]. Recently, it has been reported that glial lipid metabolism is specially connected to cerebral function, implying that lipid metabolism-derived ferroptosis might be involved in glia after ischemia onset.

3.3.1 Microglia

Evidence has suggested that fatty acids (FAs) have a close relationship with microglia [125]. Basically, FAs are present in high amounts in the cellular membranes of microglia, which can further be metabolized into bioactive derivatives to mediate the localization and function of proteins and downstream signaling pathways [126]. Microglia can also express various lipid-sensitive receptors, including toll-like receptors, triggering receptors expressed on myeloid cells 2 and receptors for FA derivatives [127]. In addition, it has been shown that FAs can be stored within lipid droplets (LDs) in microglia to regulate the neuroinflammatory response [128]. Saturated, monounsaturated, and polyunsaturated fatty acids (PUFA) are the three main families of FAs, all of which have shown strong effects on microglial cells. Microglia-activated phorbol myristate acetate treatment induces iron accumulation and leads to iron-dependent lipid peroxidation [129]. Moreover, inducible nitric oxide synthase is greatly increased to inhibit ferroptosis by decreasing arachidinate 15-lipoxygenase activity in microglia [130]. These studies indicate that microglia activation may be involved in ferroptosis via inducing lipid peroxidation.

ACSL4 is an important isozyme for the metabolism of PUFA, which results in ferroptosis-mediated cerebral damage at ischemia onset when activated [25, 131]. Cui et al. [25] reported that the ACSL4 level is inhibited in the early phase of IS through hypoxia-inducible factor 1 alpha. They demonstrated that ACSL4 aggravated neuronal death by exacerbating lipid peroxidation-derived ferroptosis and promoting inflammatory cytokine production in microglia simultaneously [25]. Hence, inhibiting microglial lipid peroxidation and inflammatory-induced ferroptosis may be therapeutic targets in treating IS.

3.3.2 Astrocytes

Lipids play a fundamental role in astrocyte function, such as energy generation, membrane fluidity, and cell-to-cell signaling [132]. It has been reported that reactive astrocytes drive the death of neurons via lipoparticle secretion [133]. Acrolein, a neurotoxin that is endogenously produced by lipid peroxidation during ischemic onset, induces astrocytic inflammation through the NLR family pyrin domain-containing protein 3 inflammasome [134, 135], suggesting that astrocytes might have neurotoxic effects on neurons via lipid peroxidation during IS.

However, studies have illustrated that the crosstalk between astrocytes and neurons might protect neurons from lipid peroxidation at ischemia onset. LDs are one of the fundamental components of lipid metabolism in astrocytes, thus playing a critical role in maintaining astrocyte–neuron homeostasis [136]. Lipid particle-mediated FA transfer is involved in the deposition of astrocytic LDs after acute stroke and protects neurons from lipid peroxidation [137]. In the acute IS model, omega-3 PUFA present a high neuroprotective and anti-inflammatory potential via targeting the function of astrocytes to alleviate neuronal injury [138]. This indicates that neuron–astrocyte metabolic coupling might suppress ferroptosis via inhibiting lipid peroxidation. Therefore, the dual role of astrocytes in lipid metabolism during IS pathogenesis needs to be investigated.

3.3.3 Oligodendrocytes

Oligodendrocytes enwrap CNS axons with myelin, which has an extremely high lipid content and a peculiar lipid composition with a ratio of 2:2:1 cholesterol:phospholipid:glycolipid. As previously mentioned, oligodendrocytes and myelin injuries underlie the neurological damage in the ischemic cerebrum [139, 140]. ATP-binding cassette transporter A1 (ABCA1) has been shown to be one of the representative genes that play a key role in lipid metabolism when treated with the ferroptosis inducer [141]. Li et al. [142] reported that ABCA1 deficiency decreases myelination and oligodendrogenesis and deteriorates the prognosis after IS, indicating that oligodendrocytes might be involved in ferroptosis via lipid metabolism at ischemia onset, but further investigation is needed.

4. Glia–Neuron Crosstalk in Ferroptosis

As we previously discussed, iron can be accumulated in glial cells and neurons. Some special routes have been investigated as iron transporters between cells, which are involved in regulating ferroptosis in the CNS [85]. It has been demonstrated that proteins responsible for iron metabolism can be transported between cells through extracellular vesicles, including exosomes [143, 144]. In addition, TfR has been reported to be transported via tunneling nanotubes, which exist in both glia and neurons [145]. This indicates a promising mode of crosstalk between glial cells and neurons in the IS pathological process.

Ferroptosis is mainly induced by the inactivation of the antioxidant system, including the cystine-glutamate antiporter (system xc-)-GSH-GPX4-dependent antioxidant defense, which results in lipid hydroperoxide deposition [146]. It is known that astrocytes can provide maintenance and energetic support for neurons through mediating oxidative stress via glucose transformation into lactate and GSH production. Mounting evidence has shown that hypoxia promotes ferroptotic cell death, which is highly connected to excessive ROS deposition by mitochondrial complex III, and enhances lipid peroxidation directly via damaging mitochondrial membranes, resulting in the elevation of hydrogen peroxide levels to initiate and aggravate Fenton reactions and/or acting through the suppression of cellular antioxidant capacity. It is known that neurons receive a continuous supply of glucose and oxygen via communication with astrocytes, indicating that astrocyte–neuron crosstalk might be indispensable for neuronal defense against oxidative stress-induced ferroptosis after ischemic onset. It is known that neurons can release damaged mitochondria and transfer them to astrocytes for disposal and recycling [147]. Notably, Hayakawa et al. [148] reported that astrocytes can produce functional extracellular mitochondria to support neuronal viability under IS conditions, indicating a new potential mechanism of neuroglial crosstalk that might be involved in ferroptosis after stroke.

5. The Potential Value of Targeting Ferroptosis in the Treatment of IS

In 2017, Tuo et al. [29] first reported that ferroptosis is a form of neuronal death occurring in IS. Interestingly, three hallmarks of ferroptosis—iron accumulation, ROS deposition, and lipid peroxidation—have already been studied as targets for IS treatments for decades. However, only one drug, edaravone, which acts as a ferroptosis inhibitor, has been applied for acute IS therapy [42, 43]. Others have been found to be effective in animal and cell models of IS. Catalysis by 12/15-lipoxygenase (12/15-LOX) of the stereospecific peroxidation of esterified PUFA to promote lipid peroxides and inhibitors of 12/15-LOX have been demonstrated to significantly relieve cerebral injury after IS [149, 150, 151]. Iron chelators, such as ciclopirox, 2,2-bipyridyl, and deferoxamine (DFO), have been extensively investigated for IS treatment [152, 153, 154, 155]. DFO, one of the best-known chelators, is reported to relieve brain damage and facilitate the neuronal recovery process in MCAO models [155]. In addition, iron regulatory proteins, including transferrin and CP, have been found to alleviate functional damage in IS models. Moreover, ferrostatin-1 and liproxstatin-1 can both significantly improve neurological deficits after IS onset in vivo [29]. Selenium-containing compounds have been found to decline infarct volume in the MCAO model [11]. However, with these being preclinical studies, large-scale clinical studies are warranted to examine the use of these ferroptosis inhibitors for IS treatment.

Traditional Chinese therapies have been widely applied for clinical treatments of IS; however, a specific mechanism has yet to be uncovered. Recently, researchers have shown that some traditional Chinese herbs can attenuate IS injuries by inhibiting ferroptosis. Carvacrol can suppress ferroptosis via upregulating GPX4 expression and mitigating lipid peroxidation after IS occurrence [156]. Naotaifang was reported to improve ischemia-induced neural injuries in rats via suppressing ferroptosis through the downregulation of TfR1/DMT1 and the upregulation of SCL7A11/GPX4 pathways [157]. Guan et al. [158] reported that galangin attenuates IR injury by activating SLC7A11 and GPX4 in gerbils to inhibit ferroptosis. Moreover, Carthamin yellow, which is extracted from safflower, has been reported to protect rats against IS via the mitigation of inflammation and ferroptosis [41].

6. Conclusions

Traditionally, IS research has focused on neurons; however, this has not resulted in a high rate of positive results for IS. As functional and structural supporting cells, neuroglia have received interest and offer promising therapeutic effects for IS treatments. On the other hand, glial cells play an important role in IS-associated neuropathologies, such as inflammation, excitotoxicity, necrotic tissue removal, oxidative stress maintenance, and neurogenesis, all of which have been associated with ferroptosis [5]. Here, we reviewed the mechanisms of iron homeostasis, ROS biology, and lipid peroxidation in different glial cells and the potential effects of neuron–glia crosstalk on IS-induced ferroptosis. However, the investigation of the role of ferroptosis in glial cells after IS and the underlying molecular pathways is still in the early stages, and several unresolved questions need to be answered. First, it is known that IS can induce inflammation, during which glial activation plays an important part. Many studies have shown that ferroptosis can be accompanied by neuroinflammation; however, the mechanistic link between ferroptosis and neuroinflammation in IS is still unclear. Another issue is that the causal relationship between iron deposition and neuronal death induced by ischemia remains uncovered. The question of whether iron deposits in glia form first and then influence the function of neurons or whether iron deposits in neurons cause glial overactivation so that more iron is transferred to the neurons, inducing neuronal death, needs to be addressed. In addition, the crosstalk between ferroptosis and other types of PCD in glial cells, such as apoptosis, necrosis, and autophagy, remains unclear. Associations between ferroptosis and other cell death pathways have been reported. For example, both necroptosis and ferroptosis contribute to IS [159]. Likewise, some modulators, such as p53, affect both apoptosis and ferroptosis [160]. Whether ferroptosis promotes other types of PCD or whether they can suppress or activate each other in neuroglia requires further investigation. All these questions should be addressed in further experimental and clinical studies to fully elucidate the roles of ferroptosis in neuroglia after IS.

Abbreviations

12/15-LOX, 12/15-lipoxygenase; ABCA1, ATP-binding cassette transporter A1; ACI, Acute cerebral infarction; ACSL4, acyl-CoA synthetase long-chain family member 4; ATP, adenosine triphosphate; BBB, blood-brain-barrier; cPKCγ, conventional protein kinase cγ; CNS, central nervous system; CP, ceruloplasmin; CoQ10, coenzyme Q10; COX2, cyclooxygenase 2; DFO, deferoxamine; DMT1, divalent metal transporter 1; eNOS, endothelial nitric oxide synthase; ETC, electron transport chain; FA, fatty acids; Fpn1, ferroportin; FTH1, porcine ferritin heavy chain; GPX4, Glutathione peroxidase 4; GR, glutathione reductase; GSH, glutathione; GSSG, glutathione disulfide; Heph, hephaestin A; HIBD, hypoxic-ischemic brain damage; HIF-1, hypoxia-inducible factor-1; HNE, hydroxynonenal; HO-1, heme oxygenase 1; IS, ischemic stroke; I/R, ischemia, reperfusion; LD, lipid droplets; LOX, lipoxygenase; LPCAT3, lysophosphatidylcholine acyltransferase 3; MBP, oligodendrocyte differentiation; MCAO, middle cerebral artery occlusion and reperfusion; NCOA4, nuclear receptor coactivator 4; MDA, malondialdehyde; NNTB1, non-transferrin bound iron; NO, nitric oxide; NOX, NADPH oxidase; OGD, oxygen-glucose deprivation; OSP, oligodendrocyte specific protein; PCD, programmed cell death; PGE2, prostaglandin E2; PMU2, pumilio 2; Prdx, peroxiredoxin-1; PTRF, polymerase I and transcript release factor; PUFA, polyunsaturated fatty acids; PVT1, plasmacytoma variant 1; RBMVECs, rat brain microvascular endothelial cells; RMS, rhabdomyosarcoma; ROS, reactive oxygen species; SAM, stroke associated microglia; Se, selenium; SIRT1, sirtuin 1; SLC3A2, subunit solute carrier family 3 member 2; SLC7A11, solute carrier family 7 member 11; SOD, superoxide dismutase; STEAP3, six-transmembrane epithelial antigen of prostate 3; Tf, transferrin receptor; TfR1, transferrin receptor R1; TLR4, toll-like receptor 4; UBIAD1, UbiA prenyltransferase domain containing 1; VGVAPG, Val-Gly-Val-Ala-Pro-Gly.

Author Contributions

YJP conceived the review and provided guidance. SYX and SMN wrote the manuscript and collected materials. CLZ was for picture processing. All authors contributed to the final manuscript. All authors contributed to editorial changes in the manuscript. All authors read and approved the final manuscript.

Ethics Approval and Consent to Participate

Not applicable.

Acknowledgment

Not applicable.

Funding

This study was funded by National Natural Science Foundation of China (82174484, 81973932), the Peak academic Talent of Jiangsu Hospital of Traditional Chinese Medicine (grant number: K2021RC24), Jiangsu Graduate Practice Innovation Program (grant number: SJCX22_0755).

Conflict of Interest

The authors declare no conflict of interest.

References
[1]
GBD 2016 Stroke Collaborators. Global, regional, and national burden of stroke, 1990-2016: a systematic analysis for the Global Burden of Disease Study 2016. The Lancet. Neurology. 2019; 18: 439–458.
[2]
Feske SK. Ischemic Stroke. The American Journal of Medicine. 2021; 134: 1457–1464.
[3]
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.
[4]
Hu HJ, Song M. Disrupted Ionic Homeostasis in Ischemic Stroke and New Therapeutic Targets. Journal of Stroke and Cerebrovascular Diseases. 2017; 26: 2706–2719.
[5]
Xu S, Lu J, Shao A, Zhang JH, Zhang J. Glial Cells: Role of the Immune Response in Ischemic Stroke. Frontiers in Immunology. 2020; 11: 294.
[6]
Valdés Hernández MDC, Case T, Chappell FM, Glatz A, Makin S, Doubal F, et al. Association between Striatal Brain Iron Deposition, Microbleeds and Cognition 1 Year After a Minor Ischaemic Stroke. International Journal of Molecular Sciences. 2019; 20: 1293.
[7]
Dávalos A, Fernandez-Real JM, Ricart W, Soler S, Molins A, Planas E, et al. Iron-related damage in acute ischemic stroke. Stroke. 1994; 25: 1543–1546.
[8]
Gill D, Monori G, Tzoulaki I, Dehghan A. Iron Status and Risk of Stroke. Stroke. 2018; 49: 2815–2821.
[9]
Słomka A, Świtońska M, Żekanowska E. Hepcidin Levels Are Increased in Patients with Acute Ischemic Stroke: Preliminary Report. Journal of Stroke and Cerebrovascular Diseases. 2015; 24: 1570–1576.
[10]
Chen Y, Fan H, Wang S, Tang G, Zhai C, Shen L. Ferroptosis: A Novel Therapeutic Target for Ischemia-Reperfusion Injury. Frontiers in Cell and Developmental Biology. 2021; 9: 688605.
[11]
Alim I, Caulfield JT, Chen Y, Swarup V, Geschwind DH, Ivanova E, et al. Selenium Drives a Transcriptional Adaptive Program to Block Ferroptosis and Treat Stroke. Cell. 2019; 177: 1262–1279.e25.
[12]
Chen J, Yang L, Geng L, He J, Chen L, Sun Q, et al. Inhibition of Acyl-CoA Synthetase Long-Chain Family Member 4 Facilitates Neurological Recovery After Stroke by Regulation Ferroptosis. Frontiers in Cellular Neuroscience. 2021; 15: 632354.
[13]
Dixon SJ, Lemberg KM, Lamprecht MR, Skouta R, Zaitsev EM, Gleason CE, et al. Ferroptosis: an iron-dependent form of nonapoptotic cell death. Cell. 2012; 149: 1060–1072.
[14]
Friedmann Angeli JP, Schneider M, Proneth B, Tyurina YY, Tyurin VA, Hammond VJ, et al. Inactivation of the ferroptosis regulator Gpx4 triggers acute renal failure in mice. Nature Cell Biology. 2014; 16: 1180–1191.
[15]
Abrams RP, Carroll WL, Woerpel KA. Five-Membered Ring Peroxide Selectively Initiates Ferroptosis in Cancer Cells. ACS Chemical Biology. 2016; 11: 1305–1312.
[16]
Xu X, Lai Y, Hua ZC. Apoptosis and apoptotic body: disease message and therapeutic target potentials. Bioscience Reports. 2019; 39: BSR20180992.
[17]
Gong YN, Guy C, Olauson H, Becker JU, Yang M, Fitzgerald P, et al. ESCRT-III Acts Downstream of MLKL to Regulate Necroptotic Cell Death and Its Consequences. Cell. 2017; 169: 286–300.e16.
[18]
Parzych KR, Klionsky DJ. An overview of autophagy: morphology, mechanism, and regulation. Antioxidants & Redox Signaling. 2014; 20: 460–473.
[19]
Stockwell BR. Ferroptosis turns 10: Emerging mechanisms, physiological functions, and therapeutic applications. Cell. 2022; 185: 2401–2421.
[20]
Cui Y, Zhang Z, Zhou X, Zhao Z, Zhao R, Xu X, et al. Microglia and macrophage exhibit attenuated inflammatory response and ferroptosis resistance after RSL3 stimulation via increasing Nrf2 expression. Journal of Neuroinflammation. 2021; 18: 249.
[21]
Jhelum P, Santos-Nogueira E, Teo W, Haumont A, Lenoël I, Stys PK, et al. Ferroptosis Mediates Cuprizone-Induced Loss of Oligodendrocytes and Demyelination. The Journal of Neuroscience. 2020; 40: 9327–9341.
[22]
Guo J, Tuo QZ, Lei P. Iron, ferroptosis, and ischemic stroke. Journal of Neurochemistry. 2023. (online ahead of print)
[23]
Lu J, Xu F, Lu H. LncRNA PVT1 regulates ferroptosis through miR-214-mediated TFR1 and p53. Life Sciences. 2020; 260: 118305.
[24]
Chen C, Huang Y, Xia P, Zhang F, Li L, Wang E, et al. Long noncoding RNA Meg3 mediates ferroptosis induced by oxygen and glucose deprivation combined with hyperglycemia in rat brain microvascular endothelial cells, through modulating the p53/GPX4 axis. European Journal of Histochemistry. 2021; 65: 3224.
[25]
Cui Y, Zhang Y, Zhao X, Shao L, Liu G, Sun C, et al. ACSL4 exacerbates ischemic stroke by promoting ferroptosis-induced brain injury and neuroinflammation. Brain, Behavior, and Immunity. 2021; 93: 312–321.
[26]
Wei H, Peng Z, Chen Y, Guo J, Chen L, Shao K. cPKCγ ameliorates ischemic injury in cultured neurons exposed to oxygen glucose deprivation/reoxygenation by inhibiting ferroptosis. Neuroscience Research. 2022; 181: 95–104.
[27]
Huang Y, Liu J, He J, Hu Z, Tan F, Zhu X, et al. UBIAD1 alleviates ferroptotic neuronal death by enhancing antioxidative capacity by cooperatively restoring impaired mitochondria and Golgi apparatus upon cerebral ischemic/reperfusion insult. Cell & Bioscience. 2022; 12: 42.
[28]
Li C, Sun G, Chen B, Xu L, Ye Y, He J, et al. Nuclear receptor coactivator 4-mediated ferritinophagy contributes to cerebral ischemia-induced ferroptosis in ischemic stroke. Pharmacological Research. 2021; 174: 105933.
[29]
Tuo QZ, Lei P, Jackman KA, Li XL, Xiong H, Li XL, et al. Tau-mediated iron export prevents ferroptotic damage after ischemic stroke. Molecular Psychiatry. 2017; 22: 1520–1530.
[30]
Fang Y, Chen X, Tan Q, Zhou H, Xu J, Gu Q. Inhibiting Ferroptosis through Disrupting the NCOA4-FTH1 Interaction: A New Mechanism of Action. ACS Central Science. 2021; 7: 980–989.
[31]
Xu Y, Liu Y, Li K, Yuan D, Yang S, Zhou L, et al. COX-2/PGE2 Pathway Inhibits the Ferroptosis Induced by Cerebral Ischemia Reperfusion. Molecular Neurobiology. 2022; 59: 1619–1631.
[32]
Jin W, Zhao J, Yang E, Wang Y, Wang Q, Wu Y, et al. Neuronal STAT3/HIF-1α/PTRF axis-mediated bioenergetic disturbance exacerbates cerebral ischemia-reperfusion injury via PLA2G4A. Theranostics. 2022; 12: 3196–3216.
[33]
Liu Q, Liu Y, Li Y, Hong Z, Li S, Liu C. PUM2 aggravates the neuroinflammation and brain damage induced by ischemia-reperfusion through the SLC7A11-dependent inhibition of ferroptosis via suppressing the SIRT1. Molecular and Cellular Biochemistry. 2023; 478: 609–620.
[34]
Zhu K, Zhu X, Sun S, Yang W, Liu S, Tang Z, et al. Inhibition of TLR4 prevents hippocampal hypoxic-ischemic injury by regulating ferroptosis in neonatal rats. Experimental Neurology. 2021; 345: 113828.
[35]
Mao R, Liu H. Depletion of mmu_circ_0001751 (circular RNA Carm1) protects against acute cerebral infarction injuries by binding with microRNA-3098-3p to regulate acyl-CoA synthetase long-chain family member 4. Bioengineered. 2022; 13: 4063–4075.
[36]
Guo XW, Lu Y, Zhang H, Huang JQ, Li YW. PIEZO1 might be involved in cerebral ischemia-reperfusion injury through ferroptosis regulation: a hypothesis. Medical Hypotheses. 2021; 146: 110327.
[37]
Novgorodov SA, Voltin JR, Gooz MA, Li L, Lemasters JJ, Gudz TI. Acid sphingomyelinase promotes mitochondrial dysfunction due to glutamate-induced regulated necrosis. Journal of Lipid Research. 2018; 59: 312–329.
[38]
Crawford RR, Prescott ET, Sylvester CF, Higdon AN, Shan J, Kilberg MS, et al. Human CHAC1 Protein Degrades Glutathione, and mRNA Induction Is Regulated by the Transcription Factors ATF4 and ATF3 and a Bipartite ATF/CRE Regulatory Element. The Journal of Biological Chemistry. 2015; 290: 15878–15891.
[39]
Meng H, Wu J, Shen L, Chen G, Jin L, Yan M, et al. Microwave assisted extraction, characterization of a polysaccharide from Salvia miltiorrhiza Bunge and its antioxidant effects via ferroptosis-mediated activation of the Nrf2/HO-1 pathway. International Journal of Biological Macromolecules. 2022; 215: 398–412.
[40]
Fu C, Wu Y, Liu S, Luo C, Lu Y, Liu M, et al. Rehmannioside A improves cognitive impairment and alleviates ferroptosis via activating PI3K/AKT/Nrf2 and SLC7A11/GPX4 signaling pathway after ischemia. Journal of Ethnopharmacology. 2022; 289: 115021.
[41]
Guo H, Zhu L, Tang P, Chen D, Li Y, Li J, et al. Carthamin yellow improves cerebral ischemia reperfusion injury by attenuating inflammation and ferroptosis in rats. International Journal of Molecular Medicine. 2021; 47: 52.
[42]
Shinohara Y, Yanagihara T, Abe K, Yoshimine T, Fujinaka T, Chuma T, et al. II. Cerebral infarction/transient ischemic attack (TIA). Journal of Stroke and Cerebrovascular Diseases. 2011; 20: S31–73.
[43]
Wang Y, Liu M, Pu C. 2014 Chinese guidelines for secondary prevention of ischemic stroke and transient ischemic attack. International Journal of Stroke. 2017; 12: 302–320.
[44]
Abdul Y, Li W, Ward R, Abdelsaid M, Hafez S, Dong G, et al. Deferoxamine Treatment Prevents Post-Stroke Vasoregression and Neurovascular Unit Remodeling Leading to Improved Functional Outcomes in Type 2 Male Diabetic Rats: Role of Endothelial Ferroptosis. Translational Stroke Research. 2021; 12: 615–630.
[45]
Ma Y, Wang J, Wang Y, Yang GY. The biphasic function of microglia in ischemic stroke. Progress in Neurobiology. 2017; 157: 247–272.
[46]
Lambertsen KL, Finsen B, Clausen BH. Post-stroke inflammation-target or tool for therapy? Acta Neuropathologica. 2019; 137: 693–714.
[47]
Liu Z, Chopp M. Astrocytes, therapeutic targets for neuroprotection and neurorestoration in ischemic stroke. Progress in Neurobiology. 2016; 144: 103–120.
[48]
Jiang CT, Wu WF, Deng YH, Ge JW. Modulators of microglia activation and polarization in ischemic stroke (Review). Molecular Medicine Reports. 2020; 21: 2006–2018.
[49]
Juurlink BH. Response of glial cells to ischemia: roles of reactive oxygen species and glutathione. Neuroscience and Biobehavioral Reviews. 1997; 21: 151–166.
[50]
Gaasch JA, Lockman PR, Geldenhuys WJ, Allen DD, Van der Schyf CJ. Brain iron toxicity: differential responses of astrocytes, neurons, and endothelial cells. Neurochemical Research. 2007; 32: 1196–1208.
[51]
Benarroch EE. Brain iron homeostasis and neurodegenerative disease. Neurology. 2009; 72: 1436–1440.
[52]
Healy S, McMahon JM, FitzGerald U. Modelling iron mismanagement in neurodegenerative disease in vitro: paradigms, pitfalls, possibilities & practical considerations. Progress in Neurobiology. 2017; 158: 1–14.
[53]
Rouault TA, Zhang DL, Jeong SY. Brain iron homeostasis, the choroid plexus, and localization of iron transport proteins. Metabolic Brain Disease. 2009; 24: 673–684.
[54]
Bishop GM, Dang TN, Dringen R, Robinson SR. Accumulation of non-transferrin-bound iron by neurons, astrocytes, and microglia. Neurotoxicity Research. 2011; 19: 443–451.
[55]
Aral LA, Ergün MA, Engin AB, Börcek AÖ, Belen HB. Iron homeostasis is altered in response to hypoxia and hypothermic preconditioning in brain glial cells. Turkish Journal of Medical Sciences. 2020; 50: 2005–2016.
[56]
Zarruk JG, Berard JL, Passos dos Santos R, Kroner A, Lee J, Arosio P, et al. Expression of iron homeostasis proteins in the spinal cord in experimental autoimmune encephalomyelitis and their implications for iron accumulation. Neurobiology of Disease. 2015; 81: 93–107.
[57]
Takeda A, Devenyi A, Connor JR. Evidence for non-transferrin-mediated uptake and release of iron and manganese in glial cell cultures from hypotransferrinemic mice. Journal of Neuroscience Research. 1998; 51: 454–462.
[58]
McCarthy RC, Sosa JC, Gardeck AM, Baez AS, Lee CH, Wessling-Resnick M. Inflammation-induced iron transport and metabolism by brain microglia. The Journal of Biological Chemistry. 2018; 293: 7853–7863.
[59]
Kaur C, Ling EA. Increased expression of transferrin receptors and iron in amoeboid microglial cells in postnatal rats following an exposure to hypoxia. Neuroscience Letters. 1999; 262: 183–186.
[60]
Rathnasamy G, Ling EA, Kaur C. Hypoxia inducible factor-1α mediates iron uptake which induces inflammatory response in amoeboid microglial cells in developing periventricular white matter through MAP kinase pathway. Neuropharmacology. 2014; 77: 428–440.
[61]
Xu YX, Du F, Jiang LR, Gong J, Zhou YF, Luo QQ, et al. Effects of aspirin on expression of iron transport and storage proteins in BV-2 microglial cells. Neurochemistry International. 2015; 91: 72–77.
[62]
Gunshin H, Mackenzie B, Berger UV, Gunshin Y, Romero MF, Boron WF, et al. Cloning and characterization of a mammalian proton-coupled metal-ion transporter. Nature. 1997; 388: 482–488.
[63]
Rathore KI, Redensek A, David S. Iron homeostasis in astrocytes and microglia is differentially regulated by TNF-α and TGF-β1. Glia. 2012; 60: 738–750.
[64]
Urrutia P, Aguirre P, Esparza A, Tapia V, Mena NP, Arredondo M, et al. Inflammation alters the expression of DMT1, FPN1 and hepcidin, and it causes iron accumulation in central nervous system cells. Journal of Neurochemistry. 2013; 126: 541–549.
[65]
Mitroshina EV, Savyuk MO, Ponimaskin E, Vedunova MV. Hypoxia-Inducible Factor (HIF) in Ischemic Stroke and Neurodegenerative Disease. Frontiers in Cell and Developmental Biology. 2021; 9: 703084.
[66]
Wang GL, Jiang BH, Rue EA, Semenza GL. Hypoxia-inducible factor 1 is a basic-helix-loop-helix-PAS heterodimer regulated by cellular O2 tension. Proceedings of the National Academy of Sciences of the United States of America. 1995; 92: 5510–5514.
[67]
Qian ZM, Wu XM, Fan M, Yang L, Du F, Yung WH, et al. Divalent metal transporter 1 is a hypoxia-inducible gene. J Cell Physiol. 2011; 226: 1596-1603.
[68]
Xu M, Li Y, Meng D, Zhang D, Wang B, Xie J, et al. 6-Hydroxydopamine Induces Abnormal Iron Sequestration in BV2 Microglia by Activating Iron Regulatory Protein 1 and Inhibiting Hepcidin Release. Biomolecules. 2022; 12: 266.
[69]
Wang J, Jiang H, Xie JX. Ferroportin1 and hephaestin are involved in the nigral iron accumulation of 6-OHDA-lesioned rats. The European Journal of Neuroscience. 2007; 25: 2766–2772.
[70]
Yang L, Zhang B, Yin L, Cai B, Shan H, Zhang L, et al. Tanshinone IIA prevented brain iron dyshomeostasis in cerebral ischemic rats. Cellular Physiology and Biochemistry. 2011; 27: 23–30.
[71]
Shin JA, Kim YA, Kim HW, Kim HS, Lee KE, Kang JL, et al. Iron released from reactive microglia by noggin improves myelin repair in the ischemic brain. Neuropharmacology. 2018; 133: 202–215.
[72]
Zrzavy T, Machado-Santos J, Christine S, Baumgartner C, Weiner HL, Butovsky O, et al. Dominant role of microglial and macrophage innate immune responses in human ischemic infarcts. Brain Pathology. 2018; 28: 791–805.
[73]
Kaur C, Ling EA. Periventricular white matter damage in the hypoxic neonatal brain: role of microglial cells. Progress in Neurobiology. 2009; 87: 264–280.
[74]
Rathnasamy G, Ling EA, Kaur C. Iron and iron regulatory proteins in amoeboid microglial cells are linked to oligodendrocyte death in hypoxic neonatal rat periventricular white matter through production of proinflammatory cytokines and reactive oxygen/nitrogen species. The Journal of Neuroscience. 2011; 31: 17982–17995.
[75]
Burkhart A, Skjørringe T, Johnsen KB, Siupka P, Thomsen LB, Nielsen MS, et al. Expression of Iron-Related Proteins at the Neurovascular Unit Supports Reduction and Reoxidation of Iron for Transport Through the Blood-Brain Barrier. Molecular Neurobiology. 2016; 53: 7237–7253.
[76]
Simpson IA, Ponnuru P, Klinger ME, Myers RL, Devraj K, Coe CL, et al. A novel model for brain iron uptake: introducing the concept of regulation. Journal of Cerebral Blood Flow and Metabolism. 2015; 35: 48–57.
[77]
Hohnholt MC, Dringen R. Uptake and metabolism of iron and iron oxide nanoparticles in brain astrocytes. Biochemical Society Transactions. 2013; 41: 1588–1592.
[78]
Qian ZM, Liao QK, To Y, Ke Y, Tsoi YK, Wang GF, et al. Transferrin-bound and transferrin free iron uptake by cultured rat astrocytes. Cellular and Molecular Biology. 2000; 46: 541–548.
[79]
Lane DJR, Robinson SR, Czerwinska H, Bishop GM, Lawen A. Two routes of iron accumulation in astrocytes: ascorbate-dependent ferrous iron uptake via the divalent metal transporter (DMT1) plus an independent route for ferric iron. The Biochemical Journal. 2010; 432: 123–132.
[80]
Kim Y, Park J, Choi YK. The Role of Astrocytes in the Central Nervous System Focused on BK Channel and Heme Oxygenase Metabolites: A Review. Antioxidants. 2019; 8: 121.
[81]
Patel BN, David S. A novel glycosylphosphatidylinositol-anchored form of ceruloplasmin is expressed by mammalian astrocytes. The Journal of Biological Chemistry. 1997; 272: 20185–20190.
[82]
Hahn P, Qian Y, Dentchev T, Chen L, Beard J, Harris ZL, et al. Disruption of ceruloplasmin and hephaestin in mice causes retinal iron overload and retinal degeneration with features of age-related macular degeneration. Proceedings of the National Academy of Sciences of the United States of America. 2004; 101: 13850–13855.
[83]
Fortna RR, Watson HA, Nyquist SE. Glycosyl phosphatidylinositol-anchored ceruloplasmin is expressed by rat Sertoli cells and is concentrated in detergent-insoluble membrane fractions. Biology of Reproduction. 1999; 61: 1042–1049.
[84]
Lu LN, Qian ZM, Wu KC, Yung WH, Ke Y. Expression of Iron Transporters and Pathological Hallmarks of Parkinson’s and Alzheimer’s Diseases in the Brain of Young, Adult, and Aged Rats. Molecular Neurobiology. 2017; 54: 5213–5224.
[85]
Qian ZM, Ke Y. Brain iron transport. Biological Reviews of the Cambridge Philosophical Society. 2019; 94: 1672–1684.
[86]
Oide T, Yoshida K, Kaneko K, Ohta M, Arima K. Iron overload and antioxidative role of perivascular astrocytes in aceruloplasminemia. Neuropathology and Applied Neurobiology. 2006; 32: 170–176.
[87]
Jeong SY, David S. Age-related changes in iron homeostasis and cell death in the cerebellum of ceruloplasmin-deficient mice. The Journal of Neuroscience. 2006; 26: 9810–9819.
[88]
Kawanami T, Kato T, Daimon M, Tominaga M, Sasaki H, Maeda K, et al. Hereditary caeruloplasmin deficiency: clinicopathological study of a patient. Journal of Neurology, Neurosurgery, and Psychiatry. 1996; 61: 506–509.
[89]
Ryan F, Zarruk JG, Lößlein L, David S. Ceruloplasmin Plays a Neuroprotective Role in Cerebral Ischemia. Frontiers in Neuroscience. 2019; 12: 988.
[90]
Yang L, Fan M, Du F, Gong Q, Bi ZG, Zhu ZJ, et al. Hypoxic preconditioning increases iron transport rate in astrocytes. Biochimica et Biophysica Acta. 2012; 1822: 500–508.
[91]
Wang Y, Liu G, Hong D, Chen F, Ji X, Cao G. White matter injury in ischemic stroke. Progress in Neurobiology. 2016; 141: 45–60.
[92]
Alix JJP, Domingues AMDJ. White matter synapses: form, function, and dysfunction. Neurology. 2011; 76: 397–404.
[93]
Kim HW, Shin JA, Kim HJ, Ahn JH, Park EM. Enhanced repair processes and iron uptake by ischemic preconditioning in the brain during the recovery phase after ischemic stroke. Brain Research. 2021; 1750: 147172.
[94]
Michalski D, Keck AL, Grosche J, Martens H, Härtig W. Immunosignals of Oligodendrocyte Markers and Myelin-Associated Proteins Are Critically Affected after Experimental Stroke in Wild-Type and Alzheimer Modeling Mice of Different Ages. Frontiers in Cellular Neuroscience. 2018; 12: 23.
[95]
Orellana-Urzúa S, Rojas I, Líbano L, Rodrigo R. Pathophysiology of Ischemic Stroke: Role of Oxidative Stress. Current Pharmaceutical Design. 2020; 26: 4246–4260.
[96]
Ren JX, Li C, Yan XL, Qu Y, Yang Y, Guo ZN. Crosstalk between Oxidative Stress and Ferroptosis/Oxytosis in Ischemic Stroke: Possible Targets and Molecular Mechanisms. Oxidative Medicine and Cellular Longevity. 2021; 2021: 6643382.
[97]
Qin C, Zhou LQ, Ma XT, Hu ZW, Yang S, Chen M, et al. Dual Functions of Microglia in Ischemic Stroke. Neuroscience Bulletin. 2019; 35: 921–933.
[98]
Li C, Zhao Z, Luo Y, Ning T, Liu P, Chen Q, et al. Macrophage-Disguised Manganese Dioxide Nanoparticles for Neuroprotection by Reducing Oxidative Stress and Modulating Inflammatory Microenvironment in Acute Ischemic Stroke. Advanced Science. 2021; 8: e2101526.
[99]
Huang Z, Qian K, Chen J, Qi Y, E Y, Liang J, et al. A biomimetic zeolite-based nanoenzyme contributes to neuroprotection in the neurovascular unit after ischaemic stroke via efficient removal of zinc and ROS. Acta Biomaterialia. 2022; 144: 142–156.
[100]
Tang XN, Cairns B, Kim JY, Yenari MA. NADPH oxidase in stroke and cerebrovascular disease. Neurological Research. 2012; 34: 338–345.
[101]
Liu J, Kang R, Tang D. Signaling pathways and defense mechanisms of ferroptosis. The FEBS Journal. 2022; 289: 7038–7050.
[102]
Jin R, Yang R, Cui C, Zhang H, Cai J, Geng B, et al. Ferroptosis due to Cystathionine γ Lyase/Hydrogen Sulfide Downregulation Under High Hydrostatic Pressure Exacerbates VSMC Dysfunction. Frontiers in Cell and Developmental Biology. 2022; 10: 829316.
[103]
Tian DS, Li CY, Qin C, Murugan M, Wu LJ, Liu JL. Deficiency in the voltage-gated proton channel Hv1 increases M2 polarization of microglia and attenuates brain damage from photothrombotic ischemic stroke. Journal of Neurochemistry. 2016; 139: 96–105.
[104]
Rhee SG, Kang SW, Jeong W, Chang TS, Yang KS, Woo HA. Intracellular messenger function of hydrogen peroxide and its regulation by peroxiredoxins. Current Opinion in Cell Biology. 2005; 17: 183–189.
[105]
Chen X, Zhao Y, Luo W, Chen S, Lin F, Zhang X, et al. Celastrol induces ROS-mediated apoptosis via directly targeting peroxiredoxin-2 in gastric cancer cells. Theranostics. 2020; 10: 10290–10308.
[106]
Kim S, Lee W, Jo H, Sonn SK, Jeong SJ, Seo S, et al. The antioxidant enzyme Peroxiredoxin-1 controls stroke-associated microglia against acute ischemic stroke. Redox Biology. 2022; 54: 102347.
[107]
Tang Z, Gan Y, Liu Q, Yin JX, Liu Q, Shi J, et al. CX3CR1 deficiency suppresses activation and neurotoxicity of microglia/macrophage in experimental ischemic stroke. Journal of Neuroinflammation. 2014; 11: 26.
[108]
Jin L, Zhu Z, Hong L, Qian Z, Wang F, Mao Z. ROS-responsive 18β-glycyrrhetic acid-conjugated polymeric nanoparticles mediate neuroprotection in ischemic stroke through HMGB1 inhibition and microglia polarization regulation. Bioactive Materials. 2022; 19: 38–49.
[109]
Tzvetanov P, Nicoloff G, Rousseff R, Christova P. Increased levels of elastin-derived peptides in cerebrospinal fluid of patients with lacunar stroke. Clinical Neurology and Neurosurgery. 2008; 110: 239–244.
[110]
Nicoloff G, Tzvetanov P, Christova P, Baydanoff S. Detection of elastin derived peptides in cerebrospinal fluid of patients with first ever ischaemic stroke. Neuropeptides. 2008; 42: 277–282.
[111]
Szychowski KA, Gmiński J. The VGVAPG Peptide Regulates the Production of Nitric Oxide Synthases and Reactive Oxygen Species in Mouse Astrocyte Cells In Vitro. Neurochemical Research. 2019; 44: 1127–1137.
[112]
George PM, Steinberg GK. Novel Stroke Therapeutics: Unraveling Stroke Pathophysiology and Its Impact on Clinical Treatments. Neuron. 2015; 87: 297–309.
[113]
Li H, Xie Y, Zhang N, Yu Y, Zhang Q, Ding S. Disruption of IP₃R2-mediated Ca²⁺ signaling pathway in astrocytes ameliorates neuronal death and brain damage while reducing behavioral deficits after focal ischemic stroke. Cell Calcium. 2015; 58: 565–576.
[114]
Sayre NL, Sifuentes M, Holstein D, Cheng SY, Zhu X, Lechleiter JD. Stimulation of astrocyte fatty acid oxidation by thyroid hormone is protective against ischemic stroke-induced damage. Journal of Cerebral Blood Flow and Metabolism. 2017; 37: 514–527.
[115]
Kang YP, Mockabee-Macias A, Jiang C, Falzone A, Prieto-Farigua N, Stone E, et al. Non-canonical Glutamate-Cysteine Ligase Activity Protects against Ferroptosis. Cell Metabolism. 2021; 33: 174–189.e7.
[116]
Tekkök SB, Ye Z, Ransom BR. Excitotoxic mechanisms of ischemic injury in myelinated white matter. Journal of Cerebral Blood Flow and Metabolism. 2007; 27: 1540–1552.
[117]
Káradóttir R, Cavelier P, Bergersen LH, Attwell D. NMDA receptors are expressed in oligodendrocytes and activated in ischaemia. Nature. 2005; 438: 1162–1166.
[118]
Domercq M, Perez-Samartin A, Aparicio D, Alberdi E, Pampliega O, Matute C. P2X7 receptors mediate ischemic damage to oligodendrocytes. Glia. 2010; 58: 730–740.
[119]
Matute C, Torre I, Pérez-Cerdá F, Pérez-Samartín A, Alberdi E, Etxebarria E, et al. P2X(7) receptor blockade prevents ATP excitotoxicity in oligodendrocytes and ameliorates experimental autoimmune encephalomyelitis. The Journal of Neuroscience. 2007; 27: 9525–9533.
[120]
Orrenius S, Zhivotovsky B, Nicotera P. Regulation of cell death: the calcium-apoptosis link. Nature Reviews. Molecular Cell Biology. 2003; 4: 552–565.
[121]
Matute C, Sánchez-Gómez MV, Martínez-Millán L, Miledi R. Glutamate receptor-mediated toxicity in optic nerve oligodendrocytes. Proceedings of the National Academy of Sciences of the United States of America. 1997; 94: 8830–8835.
[122]
Xu Y, Li K, Zhao Y, Zhou L, Liu Y, Zhao J. Role of Ferroptosis in Stroke. Cellular and Molecular Neurobiology. 2023; 43: 205–222.
[123]
Nasoohi S, Simani L, Khodagholi F, Nikseresht S, Faizi M, Naderi N. Coenzyme Q10 supplementation improves acute outcomes of stroke in rats pretreated with atorvastatin. Nutritional Neuroscience. 2019; 22: 264–272.
[124]
Belousova M, Tokareva OG, Gorodetskaya E, Kalenikova EI, Medvedev OS. Intravenous Treatment With Coenzyme Q10 Improves Neurological Outcome and Reduces Infarct Volume After Transient Focal Brain Ischemia in Rats. Journal of Cardiovascular Pharmacology. 2016; 67: 103–109.
[125]
Leyrolle Q, Layé S, Nadjar A. Direct and indirect effects of lipids on microglia function. Neuroscience Letters. 2019; 708: 134348.
[126]
Rey C, Nadjar A, Joffre F, Amadieu C, Aubert A, Vaysse C, et al. Maternal n-3 polyunsaturated fatty acid dietary supply modulates microglia lipid content in the offspring. Prostaglandins, Leukotrienes, and Essential Fatty Acids. 2018; 133: 1–7.
[127]
Mauerer R, Walczak Y, Langmann T. Comprehensive mRNA profiling of lipid-related genes in microglia and macrophages using taqman arrays. Methods in Molecular Biology. 2009; 580: 187–201.
[128]
Nadjar A. Role of metabolic programming in the modulation of microglia phagocytosis by lipids. Prostaglandins, Leukotrienes, and Essential Fatty Acids. 2018; 135: 63–73.
[129]
Yoshida T, Tanaka M, Sotomatsu A, Hirai S, Okamoto K. Activated microglia cause iron-dependent lipid peroxidation in the presence of ferritin. Neuroreport. 1998; 9: 1929–1933.
[130]
Kapralov AA, Yang Q, Dar HH, Tyurina YY, Anthonymuthu TS, Kim R, et al. Redox lipid reprogramming commands susceptibility of macrophages and microglia to ferroptotic death. Nature Chemical Biology. 2020; 16: 278–290.
[131]
Li Y, Feng D, Wang Z, Zhao Y, Sun R, Tian D, et al. Ischemia-induced ACSL4 activation contributes to ferroptosis-mediated tissue injury in intestinal ischemia/reperfusion. Cell Death and Differentiation. 2019; 26: 2284–2299.
[132]
Lee JA, Hall B, Allsop J, Alqarni R, Allen SP. Lipid metabolism in astrocytic structure and function. Seminars in Cell & Developmental Biology. 2021; 112: 123–136.
[133]
Guttenplan KA, Weigel MK, Prakash P, Wijewardhane PR, Hasel P, Rufen-Blanchette U, et al. Neurotoxic reactive astrocytes induce cell death via saturated lipids. Nature. 2021; 599: 102–107.
[134]
Park JH, Choi JY, Jo C, Koh YH. Involvement of ADAM10 in acrolein-induced astrocytic inflammation. Toxicology Letters. 2020; 318: 44–49.
[135]
Moghe A, Ghare S, Lamoreau B, Mohammad M, Barve S, McClain C, et al. Molecular mechanisms of acrolein toxicity: relevance to human disease. Toxicological Sciences. 2015; 143: 242–255.
[136]
Eraso-Pichot A, Brasó-Vives M, Golbano A, Menacho C, Claro E, Galea E, et al. GSEA of mouse and human mitochondriomes reveals fatty acid oxidation in astrocytes. Glia. 2018; 66: 1724–1735.
[137]
Ioannou MS, Jackson J, Sheu SH, Chang CL, Weigel AV, Liu H, et al. Neuron-Astrocyte Metabolic Coupling Protects against Activity-Induced Fatty Acid Toxicity. Cell. 2019; 177: 1522–1535.e14.
[138]
Zendedel A, Habib P, Dang J, Lammerding L, Hoffmann S, Beyer C, et al. Omega-3 polyunsaturated fatty acids ameliorate neuroinflammation and mitigate ischemic stroke damage through interactions with astrocytes and microglia. Journal of Neuroimmunology. 2015; 278: 200–211.
[139]
Chen J, Sun M, Zhang X, Miao Z, Chua BHL, Hamdy RC, et al. Increased oligodendrogenesis by humanin promotes axonal remyelination and neurological recovery in hypoxic/ischemic brains. Hippocampus. 2015; 25: 62–71.
[140]
Chen Y, Tian H, Yao E, Tian Y, Zhang H, Xu L, et al. Soluble epoxide hydrolase inhibition Promotes White Matter Integrity and Long-Term Functional Recovery after chronic hypoperfusion in mice. Scientific Reports. 2017; 7: 7758.
[141]
Luo T, Gao J, Lin N, Wang J. Effects of Two Kinds of Iron Nanoparticles as Reactive Oxygen Species Inducer and Scavenger on the Transcriptomic Profiles of Two Human Leukemia Cells with Different Stemness. Nanomaterials. 2020; 10: 1951.
[142]
Li L, Li R, Zacharek A, Wang F, Landschoot-Ward J, Chopp M, et al. ABCA1/ApoE/HDL Signaling Pathway Facilitates Myelination and Oligodendrogenesis after Stroke. International Journal of Molecular Sciences. 2020; 21: 4369.
[143]
Yanatori I, Richardson DR, Dhekne HS, Toyokuni S, Kishi F. CD63 is regulated by iron via the IRE-IRP system and is important for ferritin secretion by extracellular vesicles. Blood. 2021; 138: 1490–1503.
[144]
Brown CW, Amante JJ, Chhoy P, Elaimy AL, Liu H, Zhu LJ, et al. Prominin2 Drives Ferroptosis Resistance by Stimulating Iron Export. Developmental Cell. 2019; 51: 575–586.e4.
[145]
Burtey A, Wagner M, Hodneland E, Skaftnesmo KO, Schoelermann J, Mondragon IR, et al. Intercellular transfer of transferrin receptor by a contact-, Rab8-dependent mechanism involving tunneling nanotubes. FASEB Journal. 2015; 29: 4695–4712.
[146]
Chen X, Li J, Kang R, Klionsky DJ, Tang D. Ferroptosis: machinery and regulation. Autophagy. 2021; 17: 2054–2081.
[147]
Davis CHO, Kim KY, Bushong EA, Mills EA, Boassa D, Shih T, et al. Transcellular degradation of axonal mitochondria. Proceedings of the National Academy of Sciences of the United States of America. 2014; 111: 9633–9638.
[148]
Hayakawa K, Esposito E, Wang X, Terasaki Y, Liu Y, Xing C, et al. Transfer of mitochondria from astrocytes to neurons after stroke. Nature. 2016; 535: 551–555.
[149]
Liu Y, Zheng Y, Karatas H, Wang X, Foerch C, Lo EH, et al. 12/15-Lipoxygenase Inhibition or Knockout Reduces Warfarin-Associated Hemorrhagic Transformation After Experimental Stroke. Stroke. 2017; 48: 445–451.
[150]
Rai G, Joshi N, Jung JE, Liu Y, Schultz L, Yasgar A, et al. Potent and selective inhibitors of human reticulocyte 12/15-lipoxygenase as anti-stroke therapies. Journal of Medicinal Chemistry. 2014; 57: 4035–4048.
[151]
Silva BC, de Miranda AS, Rodrigues FG, Silveira ALM, Resende GHDS, Moraes MFD, et al. The 5-lipoxygenase (5-LOX) Inhibitor Zileuton Reduces Inflammation and Infarct Size with Improvement in Neurological Outcome Following Cerebral Ischemia. Current Neurovascular Research. 2015; 12: 398–403.
[152]
Prass K, Ruscher K, Karsch M, Isaev N, Megow D, Priller J, et al. Desferrioxamine induces delayed tolerance against cerebral ischemia in vivo and in vitro. Journal of Cerebral Blood Flow and Metabolism. 2002; 22: 520–525.
[153]
Freret T, Valable S, Chazalviel L, Saulnier R, Mackenzie ET, Petit E, et al. Delayed administration of deferoxamine reduces brain damage and promotes functional recovery after transient focal cerebral ischemia in the rat. The European Journal of Neuroscience. 2006; 23: 1757–1765.
[154]
Feng H, Hu L, Zhu H, Tao L, Wu L, Zhao Q, et al. Repurposing antimycotic ciclopirox olamine as a promising anti-ischemic stroke agent. Acta Pharmaceutica Sinica. B. 2020; 10: 434–446.
[155]
Xing Y, Hua Y, Keep RF, Xi G. Effects of deferoxamine on brain injury after transient focal cerebral ischemia in rats with hyperglycemia. Brain Research. 2009; 1291: 113–121.
[156]
Guan X, Li X, Yang X, Yan J, Shi P, Ba L, et al. The neuroprotective effects of carvacrol on ischemia/reperfusion-induced hippocampal neuronal impairment by ferroptosis mitigation. Life Sciences. 2019; 235: 116795.
[157]
Lan B, Ge JW, Cheng SW, Zheng XL, Liao J, He C, et al. Extract of Naotaifang, a compound Chinese herbal medicine, protects neuron ferroptosis induced by acute cerebral ischemia in rats. Journal of Integrative Medicine. 2020; 18: 344–350.
[158]
Guan X, Li Z, Zhu S, Cheng M, Ju Y, Ren L, et al. Galangin attenuated cerebral ischemia-reperfusion injury by inhibition of ferroptosis through activating the SLC7A11/GPX4 axis in gerbils. Life Sciences. 2021; 264: 118660.
[159]
Linkermann A, Skouta R, Himmerkus N, Mulay SR, Dewitz C, De Zen F, et al. Synchronized renal tubular cell death involves ferroptosis. Proceedings of the National Academy of Sciences of the United States of America. 2014; 111: 16836–16841.
[160]
Jiang L, Kon N, Li T, Wang SJ, Su T, Hibshoosh H, et al. Ferroptosis as a p53-mediated activity during tumour suppression. Nature. 2015; 520: 57–62.

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

Share
Back to top