ArticlePDF AvailableLiterature Review

The Interplay between Ferroptosis and Neuroinflammation in Central Neurological Disorders

Authors:

Abstract

Central neurological disorders are significant contributors to morbidity, mortality, and long-term disability globally in modern society. These encompass neurodegenerative diseases, ischemic brain diseases, traumatic brain injury, epilepsy, depression, and more. The involved pathogenesis is notably intricate and diverse. Ferroptosis and neuroinflammation play pivotal roles in elucidating the causes of cognitive impairment stemming from these diseases. Given the concurrent occurrence of ferroptosis and neuroinflammation due to metabolic shifts such as iron and ROS, as well as their critical roles in central nervous disorders, the investigation into the co-regulatory mechanism of ferroptosis and neuroinflammation has emerged as a prominent area of research. This paper delves into the mechanisms of ferroptosis and neuroinflammation in central nervous disorders, along with their interrelationship. It specifically emphasizes the core molecules within the shared pathways governing ferroptosis and neuroinflammation, including SIRT1, Nrf2, NF-κB, Cox-2, iNOS/NO·, and how different immune cells and structures contribute to cognitive dysfunction through these mechanisms. Researchers’ findings suggest that ferroptosis and neuroinflammation mutually promote each other and may represent key factors in the progression of central neurological disorders. A deeper comprehension of the common pathway between cellular ferroptosis and neuroinflammation holds promise for improving symptoms and prognosis related to central neurological disorders.
Citation: Xu, Y.; Jia, B.; Li, J.; Li, Q.;
Luo, C. The Interplay between
Ferroptosis and Neuroinflammation
in Central Neurological Disorders.
Antioxidants 2024,13, 395. https://
doi.org/10.3390/antiox13040395
Academic Editor: Pamela A. Maher
Received: 11 March 2024
Revised: 23 March 2024
Accepted: 25 March 2024
Published: 26 March 2024
Copyright: © 2024 by the authors.
Licensee MDPI, Basel, Switzerland.
This article is an open access article
distributed under the terms and
conditions of the Creative Commons
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
antioxidants
Review
The Interplay between Ferroptosis and Neuroinflammation in
Central Neurological Disorders
Yejia Xu 1,2,† , Bowen Jia 1,, Jing Li 1, Qianqian Li 3 ,4, * and Chengliang Luo 1,2,3,*
1Department of Forensic Medicine, School of Basic Medicine and Biological Sciences, Soochow University,
Suzhou 215123, China
2Hebei Key Laboratory of Forensic Medicine, College of Forensic Medicine, Hebei Medical University,
Shijiazhuang 050017, China
3NHC Key Laboratory of Drug Addiction Medicine, Department of Forensic Medicine, School of Forensic
Medicine, Kunming Medical University, Kunming 650500, China
4School of Forensic Medicine, Wannan Medical College, Wuhu 241002, China
*Correspondence: bzliqian@126.com (Q.L.); clluo@suda.edu.cn (C.L.)
These authors contributed equally to this work.
Abstract: Central neurological disorders are significant contributors to morbidity, mortality, and long-
term disability globally in modern society. These encompass neurodegenerative diseases, ischemic
brain diseases, traumatic brain injury, epilepsy, depression, and more. The involved pathogenesis is
notably intricate and diverse. Ferroptosis and neuroinflammation play pivotal roles in elucidating
the causes of cognitive impairment stemming from these diseases. Given the concurrent occurrence
of ferroptosis and neuroinflammation due to metabolic shifts such as iron and ROS, as well as their
critical roles in central nervous disorders, the investigation into the co-regulatory mechanism of
ferroptosis and neuroinflammation has emerged as a prominent area of research. This paper delves
into the mechanisms of ferroptosis and neuroinflammation in central nervous disorders, along with
their interrelationship. It specifically emphasizes the core molecules within the shared pathways
governing ferroptosis and neuroinflammation, including SIRT1, Nrf2, NF-
κ
B, Cox-2, iNOS/NO
·
,
and how different immune cells and structures contribute to cognitive dysfunction through these
mechanisms. Researchers’ findings suggest that ferroptosis and neuroinflammation mutually promote
each other and may represent key factors in the progression of central neurological disorders. A
deeper comprehension of the common pathway between cellular ferroptosis and neuroinflammation
holds promise for improving symptoms and prognosis related to central neurological disorders.
Keywords: ferroptosis; neuroinflammation; central neurological disorders; blood–brain barrier
1. Introduction
Nerve cell death stands as a significant contributor to neurological disorders, diseases,
and injuries. The unexpected or programmed demise of neurons and glial cells across
various regions of the nervous system disrupts sensory, motor, cognitive, learning, and
memory functions [
1
3
]. Ferroptosis, a form of regulated cell death triggered by iron
overload, involves distinct inducible factors and metabolites [
4
]. Since its inception in
2012, research has revealed that cellular ferroptosis can occur ubiquitously throughout the
human body, including within the nervous system. Imbalances in iron metabolism prompt
nerve cells to generate high levels of reactive oxygen species via the Fenton reaction, lead-
ing to lipid peroxidation and impairment of cellular REDOX functions [
5
]. Furthermore,
ferroptosis-induced mitochondrial atrophy and functional decline impede the energy sup-
ply to neurons, disrupting electrical and chemical signal transmission [
6
]. The ferroptosis of
nerve cells within the central nervous system undermines cellular physiological functions,
with cognitive impairment emerging as a hallmark mechanism in central nervous system
diseases and injuries.
Antioxidants 2024,13, 395. https://doi.org/10.3390/antiox13040395 https://www.mdpi.com/journal/antioxidants
Antioxidants 2024,13, 395 2 of 31
Neuroinflammation represents a unique inflammatory response within the human
body, serving as the central nervous system’s defense mechanism against diverse stimuli [
7
].
This process involves the activation of microglia and astrocytes, the primary immune-active
cells, which release various cytokines and chemokines [
8
,
9
]. In the context of central
nervous system diseases and injuries, neuroinflammation is a common occurrence, leading
to the release of numerous effector proteins that disrupt neuronal excitability, thereby
exacerbating nerve function impairment [10].
The coexistence of neuroinflammation and cell ferroptosis in neuropathy has sparked
significant interest in recent years within the study of central nervous system diseases.
Common inflammatory factors released during neuroinflammation exhibit close associ-
ations with proteins that govern cellular iron metabolism and lipid metabolism [
11
]. In
light of this, researchers are delving into the potential links and shared pathways between
neuroinflammation and ferroptosis by uncovering the molecular mechanisms that underlie
diverse structural alterations within cells following central nervous system diseases and
injuries.
2. Ferroptosis and Central Neurological Disorders
2.1. Mechanism and Pathway of Ferroptosis
Ferroptosis is a newly discovered type of programmed cell death, first identified in
2012 by Drs. Brent R. Stockwell, Scott Dixon, and their research team [
4
]. It is considered an
independent mechanism of cell death because cells undergoing ferroptosis exhibit distinct
morphological, compositional, and metabolic characteristics compared to common forms
of cell death, such as apoptosis and necrosis. Morphologically, cells undergoing ferroptosis
do not display fragmented or shrunken nuclei but instead exhibit atrophied mitochondria
and a reduced number of cristae [
12
]. In the cytoplasmic component, the levels of free iron
and reactive oxygen species are elevated in these cells [13].
The cellular metabolism involved in ferroptosis encompasses various facets, including
abnormalities in iron homeostasis, lipid metabolism, and glutathione-related metabolic
changes [
4
]. These alterations in metabolism often serve as critical targets for the regulation
of ferroptosis. Pertaining to iron metabolism, cells undergoing ferroptosis exhibit modifica-
tions in the processes of iron uptake, storage, utilization, and efflux, culminating in cellular
iron overload [
14
,
15
]. Researchers have addressed the issue of elevated cellular iron levels
by regulating proteins such as transferrin receptor, ferritin, ferroportin1, and employing
iron chelators like deferriamine [
14
,
16
,
17
]. In the realm of cellular lipid metabolism, the
peroxidation of polyunsaturated fatty acids into PUFA-PLs represents a pivotal pathway
in ferroptosis [
4
]. Building upon this foundation, strategies such as inhibiting PUFAacyl-
CoA derivatives (PUFA-CoAs) and ACSL4, a key enzyme generated by lipoxygenase,
or promoting monounsaturated fatty acids (MUFAs) over PUFAs have been explored as
means to impede ferroptosis [
18
20
]. Concerning intracellular REDOX metabolism, diverse
antioxidant pathways have been identified as capable of combating the progression of
cellular ferroptosis. Apart from the well-known GPX4 classical pathway, recent discoveries
have unveiled the potential of pathways like FSP1/CoQ10, DHODH, and GCH1/BH4 in in-
hibiting intracellular peroxidation by reducing CoQ10 at the plasma membrane, converting
CoQ to panthenol, and sequestering free radicals, respectively [
21
23
]. Ongoing research
continues to enhance our understanding of the mechanisms and regulation of ferroptosis,
while investigations into its interplay with other cell death mechanisms and additional
metal elements such as copper and zinc offer novel avenues for exploration [24].
2.2. Ferroptosis in Central Neurological Disorders
The typical progression of ferroptosis has been discovered in various systemic diseases.
Over the past decade, researchers have discovered that ferroptosis plays a crucial role in
many central nervous system diseases and injuries, including neurodegenerative diseases,
stroke, epilepsy, brain injury, and depression. In the early stages of research, significant
iron content and elevated transferrin levels were observed in the local blood vessels and
Antioxidants 2024,13, 395 3 of 31
brain parenchyma of ischemic stroke [
25
]. With the recognition of ferroptosis, studies
have found that ischemic stroke regulation of ACSL4 can change the pathologic severity
through the pathway of ferroptosis [
26
]. Alzheimer’s disease is characterized by the
deposition of
β
-amyloid protein (A
β
) plaques and neurofibrillary tangles (NFTs) in the
brain, which are key pathological features. There is evidence suggesting that ferroptosis
in the Alzheimer’s disease (AD) brain exhibits a unique manifestation: the abnormal
elevation of xCT exacerbates neuroexcitatory toxicity and amyloid-beta (A
β
) accumulation,
ultimately impairing the neural function in individuals with AD [
27
]. Parkinson’s disease
is marked by symptoms such as resting tremors resulting from the degeneration of neurons
in the substantia nigra of the midbrain. Research indicates that dopamine in the brain could
potentially have a positive regulatory effect on GPX4. Therefore, the depletion of dopamine
in the brain of individuals with Parkinson’s disease (PD) may lead to ferroptosis due to the
loss of GPX4 [
28
]. Furthermore, in epileptic brains, astrocytes modulate xCT/GSH/GPX4
via chemokines, leading to neuronal ferroptosis, possibly associated with synchronous
brain firing processes [
29
]. Traumatic brain injury (TBI) is a major cause of death and
disability in developed countries. The expression of TFR, FPN, and GPX4 exhibited varying
degrees of increase at 6 h, 1 day, 3 days, 7 days, and 3 weeks in the brain tissue of the injured
area. Additionally, it has been observed that ferroptosis-specific inhibitors can effectively
inhibit the extent of secondary damage following TBI [30].
The discovery of the various diseases mentioned above highlights that dysregulation
of iron metabolism, lipid metabolism, and abnormalities in GPX4-related pathways in the
affected areas have become common features of brain neurological diseases. Ferroptosis
plays a significant role in the pathogenesis of central nervous system diseases, and inter-
ventions targeting ferroptosis have become crucial strategies for combating these diseases.
By focusing on modulating ferroptosis and related pathways, researchers aim to develop
novel therapeutic approaches to address the underlying mechanisms of these neurological
disorders and potentially improve patient outcomes.
2.3. Cognitive Dysfunction Caused by Ferroptosis
The cognitive function of mammals relies on the integrity of structures such as the
cerebral cortex, hippocampus, and basal ganglia. Cognitive dysfunction is a primary
symptom of central nervous system diseases and serves as a key indicator for assessing
therapeutic effects and evaluating prognosis in numerous conditions. Neuronal damage in
various brain regions stands as the fundamental cause of cognitive impairment. Prior to
the definition of ferroptosis, numerous studies had revealed a distinct spatial and temporal
correlation between iron overload and cognitive dysfunction in the brain [
31
]. In addition
to discovering in mouse models that a high-iron diet during the neonatal period led to
cognitive impairment in adulthood, researchers also observed that iron overload in various
regions of the human brain worsened cognitive dysfunction in neurological diseases [
32
].
The inhibition of elevated iron levels in certain neurodegenerative disease models through
iron chelators has proven effective in reducing cognitive impairment [31].
In recent years, ferroptosis has been identified as prevalent in a range of central nervous
system diseases, causing damage to nerve cells and ultimately leading to neuronal death
and cognitive dysfunction. Mei and Liu’s research team illustrated that Nrf2-dependent
cortical neuronal ferroptosis induces severe cognitive impairment in mice, as evidenced by
Morris water maze tests, Rasin scores, and electroencephalograms of epileptic mice [
33
].
Studies have indicated a positive correlation between the extent of cognitive impairment
resulting from traumatic brain injury, sepsis-associated epilepsy, and epilepsy and the
degree of ferroptosis in hippocampal neurons [
34
36
]. Parkinson’s disease stems from
the ferroptosis of neurons in the substantia nigra compacta (SNpc), leading to motor
dysfunction like static tremors, as well as subcortical cognitive deficits such as executive
dysfunction and attention deficit [
37
]. The inhibition of ferroptosis to ameliorate cognitive
impairment in central nervous system diseases represents a current focal point in the
treatment of diverse conditions.
Antioxidants 2024,13, 395 4 of 31
3. Neuroinflammation in Central Neurological Disorders
Neuroinflammation is essentially an immune response triggered by physiological
abnormalities or pathological conditions of the central nervous system. It is observed
in nearly all cases of central nervous dysfunction, including aging, infectious brain dis-
eases, autoimmune encephalopathy, neurodegenerative diseases, ischemic stroke, and
brain injury [
38
43
]. The primary effector cell phenotypes of neuroinflammation are two
centrally located innate immune cells: microglia and astrocytes [
44
]. Furthermore, when
the structure and function of the blood–brain barrier, the centerpiece of immune privilege in
defending the brain, is compromised, peripheral immune cells (such as neutrophils, mono-
cytes/macrophages, lymphocytes, etc.) invade the brain and contribute to the amplification
of neuroinflammation [45].
The initiation of neuroinflammation is typically accompanied by the injury of en-
dogenous nerve cells or the invasion of exogenous infectious agents and toxins. Similar
to immune processes in other systems, damaged or dead cells and invading infectious
agents in the brain release damage-associated molecular patterns (DAMPs) and pathogen-
associated molecular patterns (PAMPs), respectively [
46
,
47
]. Microglia and astrocytes
in the brain are activated by receptors to become reactive microglia and reactive astro-
cytes [
48
,
49
]. These activated innate immune cells secrete a variety of cytokines in response
to specific phenotypic changes, including interleukins, interferons, tumor necrosis factors,
chemokines, and reactive oxygen species, which contribute to the amplification of neu-
roinflammation [50]. When pro-inflammatory cytokines take precedence, they trigger cell
death, affecting neurons, glial cells, and endothelial cells. Destruction of the endothelial
cells of the blood–brain barrier leads to the invasion of peripheral immune cells into the
brain at different stages of the disease [
51
]. These infiltrating peripheral cells further disrupt
the integrity of the brain structure, harm neurons, and worsen the organic brain damage
caused by neuroinflammation [52].
Neuroinflammation causes damage to neurons in the affected region, including axonal
degeneration, abnormal energy metabolism, destruction of synaptic structures, and even
neuronal death [
44
]. Additionally, under pathological conditions, astrocytes and microglia
lose their original physiological functions, such as neurotrophic function and maintenance
of synaptic plasticity [
53
]. The reduced level of brain-derived neurotrophic factor (BDNF)
in the brain inhibits the repair of neural synapses [
54
]. These consequences lead to the dis-
ruption of nerve impulse transmission, ultimately resulting in impaired cognitive function
and a series of brain functional abnormalities.
4. The Link between Neuroinflammation and Ferroptosis
Recent studies focusing on the brain and spinal cord have illuminated the intercon-
nected nature of cell ferroptosis and neuroinflammation in the realm of nerve injury and
degeneration. Intracellular oxidative stress stemming from metabolic irregularities dur-
ing ferroptosis, coupled with disruptions in mitochondrial structure and function, can
exacerbate neuroinflammation [
55
57
]. The altered phenotypes of immune cells during
inflammation and the release of pro-inflammatory factors can trigger ferroptosis in both
immune cells themselves and neighboring neurons [
58
]. Thus, exploring the shared molec-
ular communication processes between cellular ferroptosis and neuroinflammation offers
valuable insights into the pathogenesis of diverse central nervous system disorders, identi-
fies potential therapeutic targets, and enhances the management of cognitive, sensory, and
motor symptoms associated with these conditions. This chapter delves into the converging
pathways of ferroptosis and neuroinflammation across various central nervous system
diseases, examining cell phenotypes, specific structures, and microscopic molecules to
unveil commonalities and potential treatment avenues.
Antioxidants 2024,13, 395 5 of 31
4.1. Molecular Mechanisms and Signaling Pathways
4.1.1. SIRT1
Silent information regulator 1 (SIRT1) is a class III histone deacetylase that is widely
expressed in mammals. Recent research has convincingly demonstrated its close associa-
tion with cellular energy metabolism, inflammation, and various programmed cell death
pathways. The biological activity of SIRT1 requires the presence of NAD+ and responds
to changes in energy status by regulating the intracellular NAD+/NADH ratio [
59
,
60
].
Due to its exceptional performance in the body’s defense mechanisms and death modes
such as oxidative stress, inflammation, apoptosis, and ferroptosis, the role of SIRT1 in the
pathological process of central nervous system disease and injury is worth inferring.
The expression of SIRT1 is intricately linked to AMP-activated protein kinase (AMPK).
Activation of AMPK can enhance SIRT1 activity in a NAD+-dependent manner, thereby
inhibiting downstream inflammatory pathways and ferroptosis pathways [
61
,
62
]. More-
over, SIRT1 has the ability to elevate AMPK levels through acetylation modification [
63
].
Consequently, the reciprocal activation and regulation of AMPK and SIRT1 may establish
a positive feedback loop within cells, serving as a mechanism for the body to counteract
physiological dysfunction. Additionally, both AMPK and SIRT1 possess the capability to
modulate gene expression by co-regulating the activity of transcription factors, including
the mitochondrial functional protein PGC-1
α
[
64
,
65
]. PGC-1
α
is likely a pivotal target
of SIRT1 due to its transcriptional regulation, with the associated transcription factors
encompassing PPARα, PPARγ, FoxO1, Nrf1, Nrf2, etc. [6669].
Firstly, the mechanism by which SIRT1 repairs the damaged nervous system involves
the participation of Nrf2. Specifically, studies have indicated that following traumatic
brain injury, SIRT1 positively regulates Nrf2 concentration, inducing the expression of
Prxs (an ROS management system) and ultimately inhibiting p38 MAPK (Figure 1). This
modulation not only reduces acute neurotrauma after brain injury but also improves
cognitive function and long-term prognosis [
70
72
]. P38 is a serine/threonine protein
kinase that plays a crucial role in cell apoptosis/inflammatory response and the nervous
system [
73
,
74
]. Additionally, it has been demonstrated that two SIRT1 agonists, astaxanthin
and berberine, promote cognitive and memory recovery in mice through the SIRT1/p38
MAPK pathway [
70
,
75
]. Furthermore, the SIRT1/Nrf2 pathway mitigates ferroptosis after
brain injury by activating GPX4, a key factor in ferroptosis [
76
,
77
]. Upregulation of SIRT1
can reverse the decrease in GPX4 content in the hippocampus after ischemic hypoxic injury,
which is an important treatment target for HIBI [
76
]. Surprisingly, in some inflammatory
model diseases, intervention-induced activation of the SIRT1/Nrf2/GPX4 pathway has
been shown to alleviate inflammatory factors, suggesting a possible role of this pathway in
neuroinflammation [7880].
Secondly, SIRT1 is also involved in a cascade relationship with the important human
tumor suppressor gene p53, impacting SLC7A11, which increases intracellular and ex-
tracellular glutamate and cystine exchange, indirectly influencing GPX4’s inhibition of
ferroptosis (Figure 1). Knocking out SIRT1 and using ferrostatin-1, a specific inhibitor of
ferroptosis, have demonstrated similar inhibitory effects in the LPS-induced ferroptosis
model [
81
]. Studies on the acetylation levels of SIRT1 and P53 K382 after Baicalein treatment
have shown opposite trends [
82
], indicating that upregulation of SIRT1 may enhance the
expression of SLC7A11 by inhibiting the p53 gene. In brain injury resulting from cerebral
ischemia-reperfusion, a type of RNA-binding protein called pumilio 2 has been found to
decrease the expression of SLC7A11 by inhibiting SIRT1, leading to neuroinflammation
and ferroptosis in I/R [
83
]. It can be inferred that the SIRT1/p53/SLC7A11/GPX4 pathway
may play a key role in inducing a series of ferroptosis and neuroinflammatory effects in the
pathological processes of certain central nervous system diseases (Figure 1). Notably, due
to the distinct role of p53 in the cell cycle, the involvement of p53 in ferroptosis directly
inhibited by GPX4 differs from ferroptosis directly inhibited by SLC7A11 [84,85].
Antioxidants 2024,13, 395 6 of 31
Antioxidants 2024, 13, x FOR PEER REVIEW 6 of 32
RNA-binding protein called pumilio 2 has been found to decrease the expression of SLC7A11
by inhibiting SIRT1, leading to neuroinflammation and ferroptosis in I/R [83]. It can be inferred
that the SIRT1/p53/SLC7A11/GPX4 pathway may play a key role in inducing a series of fer-
roptosis and neuroinflammatory effects in the pathological processes of certain central nerv-
ous system diseases (Figure 1). Notably, due to the distinct role of p53 in the cell cycle, the
involvement of p53 in ferroptosis directly inhibited by GPX4 differs from ferroptosis directly
inhibited by SLC7A11 [84,85].
Figure 1. A schematic summary of the link between inflammation and ferroptosis in nerve cells.
SIRT1, Nrf2, and NF-κB intricately regulate one another, thereby influencing the magnitude of neu-
roinflammation and ferroptosis. SIRT1 and AMPK mutually boost each other’s expression, has-
tening the dissociation of Nrf2 and Keap1 through the upregulation of PGC-1α. Subsequently, Nrf2
translocates into organelles and binds to ARE, thereby orchestrating the transcriptional regulation
of various genes. Nrf2 can impede NF-κB activity via multiple pathways including the Nrf2/HO-1
pathway, STING/TBK-1 pathway, and Prxs/P38MAPK pathway. NF-κB, a pivotal transcription fac-
tor in inflammation regulation, not only triggers the NLRP3 complex for proinflammatory cytokine
release but also suppresses SIRT1 expression. Furthermore, SIRT1 inhibits p53 gene expression, in-
directly thwarting cellular ferroptosis through the GPX4 pathway. By modulating Prxs and
Fth/FPN, Nrf2 governs cellular oxidative stress and labile iron pools, thereby suppressing ferropto-
sis. NF-κB also stimulates Cox-2 and iNOS production, influencing neuroinflammation and ferrop-
tosis occurrence by impacting iron/ROS levels and inflammatory cytokines. Abbreviations: AA, ar-
achidonic acid; AMPK, adenosine 5’-monophosphate (AMP)-activated protein kinase; ARE, antiox-
idant responseelement; FPN, ferroportin; Fth, ferritin heavy; GPX4, glutathione peroxidase 4; GSH,
glutathione; GSSG, oxidized glutathione; HMGB1, high mobility group protein B1; HO-1, heme ox-
ygenase 1; LPS, lipopolysaccharide; MAPK, mitogen-activated protein kinase; MyD88, Myeloid Dif-
ferentiation Factor 88; NF-κB, nuclear factor-kappaB; Nrf2, nuclear factor erythroid-2-related factor
2; PGs, prostaglandins; PGC-1α, PPAR-γ Coactivator 1 alpha; ROS, reactive oxygen species; SIRT1,
silent information regulator 1; STING, stimulator of interferon genes; TBK-1, TANK-binding kinase
1; TLR4, Toll-like receptor 4. Solid arrows indicate facilitation of expression or relocation. Dashed
arrows suggest potential facilitation of expression but with unclear mechanisms. Inhibition arrows
signify suppression of expression. Dashed inhibition arrows indicate potential suppression of ex-
pression but with unclear mechanisms. This figure was created with BioRender.com.
Figure 1. A schematic summary of the link between inflammation and ferroptosis in nerve cells.
SIRT1, Nrf2, and NF-
κ
B intricately regulate one another, thereby influencing the magnitude of neu-
roinflammation and ferroptosis. SIRT1 and AMPK mutually boost each other’s expression, hastening
the dissociation of Nrf2 and Keap1 through the upregulation of PGC-1
α
. Subsequently, Nrf2 translo-
cates into organelles and binds to ARE, thereby orchestrating the transcriptional regulation of various
genes. Nrf2 can impede NF-
κ
B activity via multiple pathways including the Nrf2/HO-1 pathway,
STING/TBK-1 pathway, and Prxs/P38MAPK pathway. NF-
κ
B, a pivotal transcription factor in
inflammation regulation, not only triggers the NLRP3 complex for proinflammatory cytokine release
but also suppresses SIRT1 expression. Furthermore, SIRT1 inhibits p53 gene expression, indirectly
thwarting cellular ferroptosis through the GPX4 pathway. By modulating Prxs and Fth/FPN, Nrf2
governs cellular oxidative stress and labile iron pools, thereby suppressing ferroptosis. NF-
κ
B also
stimulates Cox-2 and iNOS production, influencing neuroinflammation and ferroptosis occurrence
by impacting iron/ROS levels and inflammatory cytokines. Abbreviations: AA, arachidonic acid;
AMPK, adenosine 5’-monophosphate (AMP)-activated protein kinase; ARE, antioxidant respon-
seelement; FPN, ferroportin; Fth, ferritin heavy; GPX4, glutathione peroxidase 4; GSH, glutathione;
GSSG, oxidized glutathione; HMGB1, high mobility group protein B1; HO-1, heme oxygenase 1;
LPS, lipopolysaccharide; MAPK, mitogen-activated protein kinase; MyD88, Myeloid Differentia-
tion Factor 88; NF-
κ
B, nuclear factor-kappaB; Nrf2, nuclear factor erythroid-2-related factor 2; PGs,
prostaglandins; PGC-1
α
, PPAR-
γ
Coactivator 1 alpha; ROS, reactive oxygen species; SIRT1, silent
information regulator 1; STING, stimulator of interferon genes; TBK-1, TANK-binding kinase 1; TLR4,
Toll-like receptor 4. Solid arrows indicate facilitation of expression or relocation. Dashed arrows
suggest potential facilitation of expression but with unclear mechanisms. Inhibition arrows signify
suppression of expression. Dashed inhibition arrows indicate potential suppression of expression but
with unclear mechanisms. This figure was created with BioRender.com.
Furthermore, SIRT1 has been found to inhibit ferroptosis in nerve cells through a non-
GPX4-dependent ferroptosis protection pathway. In experiments with HT22 cells
in vivo
and
in vitro
in Alzheimer’s mice, Jiatong Zhang and Wei Li’s team demonstrated that the
upregulation of SIRT1 in AD inhibits ferroptosis via the FSP1 axis [
86
]. Similar results
were observed in a study on subarachnoid hemorrhage [
87
]. Moreover, a test of methyl
Antioxidants 2024,13, 395 7 of 31
naphthoquinone-4 (MK-4) suggests that SIRT1 may also protect neurons from ferroptosis
in subarachnoid hemorrhage by increasing DHODH levels [88].
Finally, SIRT1 plays a crucial role in attenuating various neuroinflammatory cascades
by downregulating nuclear factor-
κ
B (NF-
κ
B), a central molecule in inflammation
(Figure 1)
.
The regulatory function of SIRT1 in inflammation can involve HMGB1 and Toll-like re-
ceptors (TLR4) [
89
]. In cases of brain injury, SIRT1 is involved in reducing NF-
κ
B levels
through the HMGB1/TLR4 pathway [
90
] and inhibiting the expression of NLRP3 inflamma-
somes [
61
]. The downstream inflammatory pathway associated with NF-
κ
B activation can
result in the upregulation of various pro-inflammatory factors and cell phenotype transfor-
mations, such as microglia and astrocyte activation [
59
], M1 polarization of microglia [
91
],
and an increase in TNF-
α
, IL-1
β
, IL-6, and other inflammatory mediators [
59
,
91
]. Numer-
ous biomolecules activated by SIRT1 work to reduce NF-
κ
B expression, thereby helping to
alleviate neuroinflammation in the brain [59,60,89,92].
It is evident that the activation of silent information regulator 1 can trigger both
ferroptosis and inflammation in cells, usually through upstream regulatory means. As such,
upregulation of SIRT1 has emerged as a potential therapeutic target for the treatment of
neurological disorders and cognitive impairment in the brain. Resveratrol, a classic SIRT1
activator, has been widely used to reduce inflammation and cell death in diseases
[89,9395]
.
However, there is still much to learn about the mechanism by which SIRT1 regulates
ferroptosis in nerve cells and inhibits the release of pro-inflammatory factors. For instance,
the specific intermolecular processes by which SIRT1 regulates FSP1 and DHODH are not
yet fully understood and disease models targeting these two atypical ferroptosis defense
mechanisms are limited. Furthermore, the SIRT1/p53 pathway has different direct effects
on GPX4 and SLC7A11, and the feedback relationship between p53, GPX4, and SLC7A11
requires further investigation.
4.1.2. Nrf2
The nuclear factor erythroid 2-related factor 2 (Nrf2) is a DNA-binding protein that
has its origins in the study of red blood cell production. While initial research on Nrf2
focused on its involvement in regulating globin gene expression, subsequent studies
have highlighted its crucial role in combating oxidative stress [
96
]. Nrf2 accomplishes
this by activating a variety of genes through binding to antioxidant response elements
(AREs), particularly promoting the expression of heme oxygenase-1 (HO-1) (Figure 1). This
activation of downstream pathways leads to an enhanced antioxidant response in different
types of cells.
The regulation of Nrf2 activity encompasses keap1-dependent and keap1-independent
pathways, with the keap1-Cul3-Rbx1 pathway playing a pivotal role. Upon exposure
to oxidative stress, the dissociation of keap1 from Nrf2 results in enhanced stability of
Nrf2 and its subsequent translocation into the nucleus to bind ARE [
96
]. In the context of
cerebral ischemia-reperfusion injury, the NLRP3 inflammasome can trigger downstream
antioxidant responses and influence ferroptosis by upregulating Nrf2 induced by keap1 as
a defensive reflex (Figure 1).
There are two primary forms of keap1-independent regulation. The first is the
PI3K/Akt-dependent pathway, which is activated in response to oxidative stress [
97
]. In
mouse HT22 cells, blocking the PI3K/Akt pathway hinders the antioxidant effect induced
by Nrf2, ultimately exacerbating neuroinflammation [
98
]. In cases of traumatic brain injury
(TBI), activation of the TrkB receptor upregulates Nrf2 through the PI3K/Akt pathway,
thereby reducing ferroptosis and TBI-related neuroinflammation. This process also partially
restores neurocognitive impairment. Consequently, the TrkB activator NAS has emerged as
an effective therapeutic target [
99
]. The second type involves the Hrd1-dependent path-
way, where Hrd1 binds to Nrf2, leading to its ubiquitination and subsequent degradation.
Consequently, this inhibits the cellular protective response [100].
Nrf2, upon translocating into the nucleus, plays a crucial role in the regulation of
more than 250 genes by binding to their specific sites. These genes are involved in various
Antioxidants 2024,13, 395 8 of 31
cellular processes such as oxidative stress, ferroptosis, autophagy, and inflammation [
101
].
In the context of ferroptosis, Nrf2 not only modulates intracellular iron, lipid, and redox
metabolism but also exerts influence on key ferroptosis inhibition pathways like GPX4
and FSP1 through intricate molecular interactions [
96
,
102
]. Reactive oxygen species (ROS)
generated during cellular stress trigger the activation of Nrf2 via the keap1-dependent
pathway [
103
]. And then, Nrf2 reduces intracellular iron by increasing the expression of
Fth and FPN, utilizing iron storage and iron effluence, respectively, and inhibits glia-driven
neuronal ferroptosis by positively regulating GPX4 and FSP1 pathways [
104
,
105
]. Netrin-1,
a binding protein associated with nerve regeneration, regulates ferroptosis after brain injury
by influencing Nrf2 through neurite inducible factor receptor UNC5B, providing a thera-
peutic direction for subsequent nerve recovery [
106
]. The peroxisome proliferator-activated
receptor PPAR
γ
pathway, which is associated with lipid metabolism, complements Nrf2
and significantly reduces ferroptosis and cognitive impairment in pathological models of
cerebral hemorrhage and cerebral ischemia [98,107,108].
In addition to its role in cell death and oxidative stress, recent research has highlighted
Nrf2’s involvement in neuroinflammation. Studies on cognitive function in offspring rats
subjected to maternal sleep deprivation have demonstrated that the ferroptosis inhibitor
liproxstatin-1 can reduce the release of inflammatory factors and the activation of microglia
via the Nrf2/HO-1 axis [
109
]. Similarly, in patients with neuroinflammation due to inter-
stitial cystitis and in models of RSL3-induced ferroptosis, Nrf2 has been observed to be
associated with pro-inflammatory cytokines and microglia [
110
,
111
]. In cases of traumatic
brain injury, Nrf2 promotes mitochondrial function recovery, enhances antioxidant capacity,
and inhibits neuroinflammation [
112
]. The mechanism by which Nrf2 regulates neuroin-
flammation is mainly dependent on its inhibition of downstream inflammatory pathways
by influencing the activation of NF-
κ
B (Figure 1). Nrf2 inhibits NF-
κ
B through various
means, including the induction of a shift in microglia from a pro-inflammatory M1 pheno-
type to an anti-inflammatory M2 phenotype via the Nrf2/STING/NF-
κ
B pathway [
113
,
114
].
The stimulator of interferon genes (STING), which serves as an immune sensor of endoge-
nous and exogenous DNA, may act as an intermediary in the NF-
κ
B induction by Nrf2
(Figure 1). While the relationship between Nrf2 and STING has been confirmed by the
STRING database, the cascading relationship between them remains to be elucidated [
115
].
The NRF2-STING axis plays a key role in regulating oxidative stress, inflammation, and
necrosis [
116
]. Activated STING can recruit TANK-binding kinase 1 (TBK1) to accelerate
the phosphorylation of NF-
κ
B [
117
]. In addition, heme oxygenase HO-1, the main regu-
latory enzyme of Nrf2, is believed to reduce NF-
κ
B activity [
118
]. Annexin A5 facilitates
phenotype transformation of microglia via the Nrf2 pathway to NF-
κ
B, counteracting the
inflammatory response, ferroptosis, and oxidative stress, thereby improving neurocognitive
dysfunction [
119
]. The effect of Nrf2 on neuroinflammation heavily relies on NF-
κ
B, but
the mechanisms by which Nrf2 regulates NF-
κ
B are varied. The association between these
two molecules remains incomplete, and the Nrf2/STING/TBK1/NF-
κ
B pathway in the
brain requires more detailed detection of protein expression and pathological experiments
for verification. In particular, the indirect relationship between Nrf2 and STING needs
further elaboration.
In summary, Nrf2, a key player in ferroptosis and oxidative stress, is significantly
involved in the connection between neuroinflammation and ferroptosis in the central
nervous system. However, our understanding of the molecular mechanisms underlying
neuroinflammation remains partial and incomplete. The downstream pathway activated
by Nrf2 within the complete neuroinflammatory signaling network should be viewed as
three-dimensional rather than linear, requiring further in-depth exploration and discussion.
4.1.3. NF-κB
Nuclear factor-
κ
B (NF-
κ
B) is a transcription factor found in all animal cell types that
plays a crucial role in immune and inflammatory responses [
120
]. As such, researchers
have focused on understanding the NF-
κ
B-related pathway, which presents a promising
Antioxidants 2024,13, 395 9 of 31
target for addressing a variety of pathological processes such as tumor progression, in-
jury, and inflammation [
121
123
]. The NF-
κ
B pathway has been implicated in regulating
neuroinflammation following central nervous system diseases, making it a particularly
important area of study [
124
]. Moreover, early research suggested, through kinase analysis,
that NF-
κ
B may play a critical role in ferroptosis [
125
]. Therefore, NF-
κ
B holds the potential
to emerge as a key focal point in the molecular mechanisms of both ferroptosis and neu-
roinflammation within the nervous system. A thorough understanding of the activation
process and mechanism of NF-
κ
B is essential for a more comprehensive prediction of the
resulting damage and identification of potential intervention targets in central nervous
system diseases.
Under normal physiological conditions, NF-
κ
B remains inactive, typically bound to
I-
κ
B-
α
in the cytoplasm. However, when there is a change in the cytoplasmic environment
or upon exposure to specific external stimuli, NF-
κ
B and I-
κ
B-
α
dissociate and translocate
into the nucleus, where they regulate the transcription of specific genes, thereby activating
NF-
κ
B [
103
] (Figure 1). In the context of neuroinflammation, common cellular triggers
that lead to NF-
κ
B activation include reactive oxygen species (ROS), damage-associated
molecular patterns (DAMPs), and lipopolysaccharides (LPSs) [103].
The generation of reactive oxygen species (ROS) is a key characteristic of cellular
ferroptosis, with the disruption of cellular iron homeostasis leading to a sudden increase
in intracellular ROS via the Fenton reaction [
12
]. In addition to instigating intracellular
lipid peroxidation and oxidative stress, ROS can also activate the NF-
κ
B pathway, thereby
triggering further inflammatory cascades [
126
]. A recent study on ferroptosis demonstrated
that reducing ROS-induced NF-
κ
B activation and inhibiting the resulting neuroinflamma-
tory response can be achieved through the use of the iron-chelating agent deferoxamine.
The study further proposed that these effects may be mediated, at least in part, by the
modulation of protein levels of p-NF-
κ
Bp65 [
126
]. This suggests that the regulation of
neuroinflammation from the standpoint of iron homeostasis is, to some extent, dependent
on NF-κB activity.
DAMPs serve as ubiquitous “warning signals” released from damaged or deceased
cells, including neurons and glial cells within the brain [
127
]. One prominent DAMP in
neuroinflammation is the high mobility group protein, HMGB1 [
128
]. In a study utilizing
ischemia-reperfusion injury as a model, it was observed that IRI led to an increase in
HMGB1 expression and upregulation of NF-
κ
B p65 in brain tissue. Interestingly, the
well-known ferroptosis inhibitor Fer-1 and its analogue Srs11-92 were able to reverse
this phenomenon [
129
]. Notably, annexin A5, an annexin protein, has been shown to
mitigate inflammation and ferroptosis following traumatic brain injury by targeting both
the HMGB1/NF-κB pathway and the Nrf2 pathway [119].
Lipopolysaccharides have the ability to directly induce neuroinflammation and are
frequently employed in constructing neuroinflammation models for numerous animal
and cell culture experiments. Research has illustrated that lipopolysaccharide-induced
neuroinflammation exerts its effects through the NF-
κ
B pathway in cells such as microglia
and neurons [130132].
Within nerve cells, DAMPs and LPS function as ligands that bind to Toll-like receptor
4 (TLR4), initiating the activation of intracellular signaling pathways [
132
134
]. TLR4
triggers the activation of NF-
κ
B via a splice protein known as myeloid differentiation
primary response protein 88 (MyD88) [131,134] (Figure 1).
Upon entering the nucleus, NF-
κ
B activates the NLRP3 inflammasome and caspase-
1, consecutively. Subsequently, the activated caspase-1 converts the precursors of IL-1
β
and IL-18 into their active conformation [
126
]. The activation of the NF-
κ
B pathway in
innate immune cells within the brain ultimately elicits a shift towards a pro-inflammatory
phenotype, ultimately resulting in the release of pro-inflammatory cytokines including
TNF-
α
, IL-6, and IL-1
β
. These cytokines can exacerbate neuroinflammation [
135
,
136
].
Amongst these pro-inflammatory factors, IL-6 can contribute to ROS production [
103
],
Antioxidants 2024,13, 395 10 of 31
whereas TNF-
α
facilitates p65 binding factors to positively regulate NF-
κ
B via nuclear
signaling [
137
]. This amplifying effect may propagate inflammation throughout the brain.
The Fe
2+
/ROS/NF-
κ
B inflammatory pathway serves as a crucial link between iron
metabolism and cellular inflammation, offering a potential avenue for inhibiting neu-
roinflammation by regulating iron levels. Following brain injury, the administration
of the iron-chelating agent DFO has been shown to elevate the protein level of p-NF-
κ
Bp65. This intervention not only diminishes the activation and infiltration of neutrophils
and macrophages but also reverses the shift towards a pro-inflammatory phenotype in
microglia-like cells, ultimately ameliorating cognitive dysfunction in mice [
126
]. Further-
more, the classical inhibitory pathway of ferroptosis, the GPX4 pathway, is intricately
connected to the pro-inflammatory effects of NF-
κ
B. Studies indicate that GPX4 can impede
TNF-
α
-mediated NF-
κ
B transcriptional activity [
102
]. The capacity of Saikosaponin B2 to
suppress ferroptosis and neuroinflammation via the TLR4/NF-
κ
B pathway relies on the
presence of GPX4 [138].
NF-
κ
B is indirectly modulated by SIRT1 and Nrf2, while it also influences the ex-
pression of SIRT1 and Nrf2 through its transcriptional products (Figure 1). The NF-
κ
B
transcription-induced inflammasome NLRP3 hampers SIRT1 synthesis [
103
]. Research
indicates that the keap1-dependent pathway, which culminates in Nrf2 activation, is linked
to the transcriptional activity of NF-
κ
B p65 [
139
]. The interplay among the three intracellu-
lar regulatory molecules—SIRT1, Nrf2, and NF-
κ
B—dictates the fate of neurons and glial
cells. The cellular REDOX state may dictate the expression levels of these three proteins
in response to external stimuli, with this equilibrium ultimately determining whether the
nerve cell veers towards a pro-inflammatory and ferroptotic outcome.
4.1.4. PTGS2/Cox-2
PTGS2, also known as prostaglandin-endoperoxide synthase 2, is a gene that regulates
the synthesis of induced cyclooxygenase Cox-2. This physiological regulator catalyzes
arachidonic acid (AA) to produce prostaglandins (PGs) and acts as a peroxidase
[140,141]
.
Cox-2 is present in the human central nervous system under normal conditions and plays an
important role in brain functions such as memory [
142
]. However, under pathological con-
ditions, Cox-2 contributes to various neurodegenerative diseases and nerve injuries [
142
].
Elevated Cox-2 levels in cerebral stroke, traumatic brain injury, and neurodegeneration lead
to pathological changes including blood–brain barrier breakdown and cerebral edema [
143
146
]. Furthermore, due to its role in lipid peroxidation during the catalytic process AA,
Cox-2 serves as a lipid metabolic indicator of ferroptosis [
147
]. Additionally, Cox-2 is highly
expressed in inflammation and correlates with inflammatory factors such as IL-1
β
, IL-6,
and TNF-α[148].
Cox-2 primarily functions in acute inflammation, experiencing rapid upregulation
following neuroinflammation and returning to baseline levels within a few hours [
142
].
The transcription factor NF-
κ
B upregulates Cox-2 expression at the onset of inflammation
by binding to Cox-2 promoters [
149
]. In models of LPS-induced neuroinflammation, NF-
κ
B
upregulation leads to elevated levels of Cox-2 and iNOS, initiating a cascade of acute
inflammatory processes [
150
]. Given its crucial role in inflammation, Cox-2 inhibition
has emerged as a significant therapeutic target for various inflammatory conditions [
142
].
Inhibiting Cox-2 has been demonstrated to suppress microglial activation, thereby reducing
neuroinflammation, and downregulating Cox-2 also decreases NF-
κ
B expression [
151
].
For example, in mouse models post-splenectomy, administration of the Cox-2 inhibitor
meloxicam effectively mitigated neuroinflammation during the acute phase and alleviated
cognitive dysfunction in the chronic phase [152].
Research into the impact of Cox-2 on ferroptosis extends beyond lipid peroxidation.
Bioinformatics analysis has identified PTGS2 as a central gene in ferroptosis biology, closely
associated with key players such as ACSL4 and GPX4 [
153
]. The regulation of Cox-2 is
intertwined with ACSL4 expression, as evidenced by the concurrent decrease in Cox-2 con-
tent following ACSL4 gene knockdown [
154
]. Similarly, loss of GPX4 can result in increased
Antioxidants 2024,13, 395 11 of 31
PTGS2 expression [
155
]. Moreover, Cox-2 overexpression has been shown to diminish
glutathione (GSH) levels through its interaction with GPX4 [
156
]. Consequently, Cox-2
expression can be modulated through the conventional ferroptosis inhibition pathway. It
is evident that the expression of PTGS2 can be significantly elevated under the influence
of ferroptosis-specific inducers like RSL3 and erastin, thereby impacting arachidonic acid
metabolism and eicosanoid biosynthesis [
135
,
140
,
157
]. Conversely, Cox-2 inhibition can
also influence the extent of ferroptosis. For instance, in a diffuse brain injury model, meloxi-
cam treatment increased GSH levels and decreased malondialdehyde, an end product of
ferroptosis, over a 48 h period [
143
]. Additionally, atorvastatin has been demonstrated to
mitigate ferroptosis-induced myocardial injury through the PTGS2 pathway [158].
Particularly, research on the regulation of Cox-2 has revealed the existence of two
types of non-coding RNAs (ncRNAs). One type is microRNA, a single-stranded RNA
capable of binding to the 3’-untranslated region (UTR) of specific targeted mRNA [
140
].
The other type is long non-coding RNA (lncRNA), a long-stranded RNA that interferes
with mRNA splicing and influences downstream gene expression. Importantly, lncRNAs,
miRNAs, and downstream genes can form pathways that regulate proteins, known as com-
petitive endogenous RNAs (ceRNAs) [
159
]. Based on targetscan, a database for predicting
miRNA target genes, and bioinformatics analysis, ceRNAs have demonstrated exceptional
performance in regulating PTGS2/Cox-2. In various systems, evidence has shown that
miR-26a-5p and lncRNA Gm47283/miR-706 influence ferroptosis in bronchial epithelial
cells in COPD and myocardial infarction via PTGS2, respectively [
159
,
160
]. Similarly,
lncRNA-Cox2 serves as an inflammatory mediator, initiating a cascade of downstream
inflammatory reactions by inducing transcription of NF-
κ
B [
161
]. Numerous examples also
illustrate how microRNAs and lncRNAs regulate PTGS2 to impact neuroinflammation and
ferroptosis in brain disease models. For instance, miR-202-3p negatively regulates PTGS2
to attenuate neuroinflammation in IS [
162
]. In Alzheimer ’s disease, lncMALAT1 negatively
regulates the expression of Mir-125b-mediated PTGS2, while miR-103 promotes nerve re-
covery and growth by targeting PTGS2 [
163
,
164
]. Additionally, miR-137 from exosomes in
endothelial progenitor cells can counteract the neuroprotective effect of SH-SY5Y (derived
from neuroblasts) by inhibiting Cox-2 [
165
]. In rats, miR-194-5p targeting PTGS2 leads
to nerve damage in temporal lobe epilepsy [
166
]. Studies have also found examples of
miR-212-5p targeting PTGS2 to reverse ferroptosis in neurons in controlled cortical impact
mouse models [140].
Indeed, the discovery of ceRNAs as upstream targets of PTGS2 that regulate Cox-2
protein expression presents exciting potential for microRNAs to modulate neuroinflam-
mation and ferroptosis. As such, further research into the microRNA/PTGS2 pathway at
the gene level is crucial. This will not only enhance our understanding of the interplay
between neuroinflammation and ferroptosis in the central nervous system, but also pave
the way for the development of corresponding interventional strategies for treating various
clinical diseases.
4.1.5. iNOS/NO
Inducible nitric oxide synthase (iNOS) is an enzyme primarily responsible for catalyz-
ing the synthesis of nitric oxide (NO). The metabolites of NO can generate nitroso salts,
leading to oxidative reduction metabolic disorders [
167
]. iNOS has been detected in nerve
damage caused by various factors and is closely associated with neuroinflammation and
neuropathic pain [
168
]. The expression of iNOS has also been found in diverse central
nervous system diseases, such as Alzheimer’s disease, Parkinson’s disease, ischemic brain
injury, and traumatic brain injury [
169
172
]. However, the specific mechanisms through
which iNOS influences the damage and long-term prognosis of these diseases are not
yet fully understood, as compared to other molecules. This article aims to explore its
relationship with ferroptosis and neuroinflammation in central nervous system diseases.
The neuroinflammation in the central nervous system caused by iNOS is closely linked
to microglia. During traumatic neuroinflammation, microglia-produced TNF-
α
activates
Antioxidants 2024,13, 395 12 of 31
the NF-
κ
B/iNOS pathway, leading to alterations in brain microcirculation function [
172
]
(Figure 1). This type of nerve injury is primarily characterized by damage to the blood–brain
barrier. In cerebral stroke, microglia rapidly produce iNOS upon cytokine stimulation [
150
].
Mechanistically, the regulation of iNOS mainly occurs at the transcriptional level and can
be influenced by downstream molecules in the neuroinflammatory signaling network,
consequently affecting upstream inflammatory factors. Following ischemic stroke, iNOS
production enhances the expression of CD14, thereby activating the TLR4/NF-
κ
B pathway
and inducing the generation of inflammatory factors [
171
]. In neuroinflammatory models
treated with LPS, microglia induce depression-like symptoms through the NLRP3/NF-
κ
B/iNOS signaling pathway [
173
]. In the treatment of epilepsy, miconazole alleviates
epilepsy symptoms by inhibiting both NF-
κ
B and iNOS [
174
]. Thus, the interaction between
NF-
κ
B and iNOS plays a pivotal role in the development of central nervous system diseases.
The activity of iNOS and the resulting NO production significantly impact the pheno-
type of microglia, potentially serving as a critical factor in triggering neuroinflammation.
LPS-treated microglia exhibit the M1 phenotype, accompanied by elevated levels of inflam-
matory factors and iNOS [
175
]. Studies indicate that the upregulation of iNOS/NO
and
the rise of reactive oxygen species (ROS) due to disrupted iron metabolism lead to height-
ened resistance to ferroptosis in M1-polarized microglia, promoting a global shift toward
the M1 phenotype [
176
,
177
]. This could explain the proinflammatory role of NO. Regarding
the mechanism through which iNOS inhibits ferroptosis in microglia and macrophages,
research has demonstrated that NO can independently regulate ferroptosis by suppressing
the production of 15-HpETE-PE [178,179].
However, focusing solely on the effect of iNOS on microglia is one-sided, as numerous
studies have demonstrated that iNOS and NO exert multiple effects on the ferroptosis of
various cell phenotypes. For instance, in individuals who smoke cigarettes, NO produced
by iNOS promotes ferroptosis in malignant mesothelioma (MM) cells through the pro-
duction of peroxynitrite [
180
]. Furthermore, iNOS activation leads to the upregulation of
ROS and RNS in triple-negative breast cancer cells, ultimately resulting in ferroptosis [
181
].
Interestingly, for the M1 and M2 phenotypes of macrophages, NO confers resistance to
ferroptosis [
182
]. Within the current research landscape, NO has exhibited a protective
effect against ferroptosis in certain immune cells, while demonstrating a sensitizing effect
in other cells. Unraveling the underlying reasons behind these contrasting effects warrants
further investigation.
Furthermore, contrary to the widely accepted notion of iNOS exerting detrimental
effects on central nervous system diseases, emerging evidence suggests that iNOS may
confer long-term protective effects on nerves [
183
]. This phenomenon could be attributed
to the bidirectional impact of NO on lipid peroxidation. In the short term, its generation of
peroxynitrite can mediate post-traumatic neuroinflammation and oxidative stress. How-
ever, owing to its ability to interact with other free radicals, NO also possesses unique
antioxidant properties [
183
], thereby safeguarding nerves in the long run. Consequently, a
comprehensive exploration of iNOS and NO warrants a closer examination of the temporal
dynamics of disease and the intricate interplay between NO and other free radicals.
4.2. Cell Phenotype and Specific Structure
4.2.1. Blood–Brain Barrier
The blood–brain barrier (BBB) plays a crucial role in facilitating communication be-
tween the brain and external substances. Any structural or functional abnormalities in the
BBB can result in brain injury and cognitive dysfunction. Notably, numerous neurodegen-
erative diseases are marked by the impairment of the blood–brain barrier, such as stroke,
traumatic brain injury (TBI), Alzheimer’s disease (AD), and Parkinson’s disease (PD) [
184
].
Therefore, it is imperative to gain a deeper understanding of the blood–brain barrier’s
involvement in the pathological processes of central nervous system diseases.
The blood–brain barrier is primarily composed of brain microvascular endothelial
cells, pericytes, basement membrane, astrocyte terminal feet, and intercellular tight junc-
Antioxidants 2024,13, 395 13 of 31
tion proteins, such as claudin-5 and ZO-1(Figure 2). Among these components, brain
microvascular endothelial cells (BMECs) and tight junction structures (TJs) play the most
crucial roles in determining the blood–brain barrier’s functionality [
185
,
186
]. Notably,
BMECs lack transcellular fenestrae, which restricts the transcellular exchange pathway
for substances. Additionally, the tight connection between BMECs minimizes intercellular
substance exchange. When the blood–brain barrier is compromised, it often results in the
infiltration of inflammatory mediators and toxic substances from the bloodstream into the
brain [
187
]. This can occur through the cell death of brain microvascular endothelial cells,
disruption of the tight junction structure, or a combination of both mechanisms, leading to
a cascade of brain injuries.
Antioxidants 2024, 13, x FOR PEER REVIEW 13 of 32
BBB can result in brain injury and cognitive dysfunction. Notably, numerous neurodegen-
erative diseases are marked by the impairment of the blood–brain barrier, such as stroke,
traumatic brain injury (TBI), Alzheimer’s disease (AD), and Parkinson’s disease (PD)
[184]. Therefore, it is imperative to gain a deeper understanding of the blood–brain bar-
rier’s involvement in the pathological processes of central nervous system diseases.
The blood–brain barrier is primarily composed of brain microvascular endothelial
cells, pericytes, basement membrane, astrocyte terminal feet, and intercellular tight junc-
tion proteins, such as claudin-5 and ZO-1(Figure 2). Among these components, brain mi-
crovascular endothelial cells (BMECs) and tight junction structures (TJs) play the most
crucial roles in determining the blood–brain barrier’s functionality [185,186]. Notably,
BMECs lack transcellular fenestrae, which restricts the transcellular exchange pathway for
substances. Additionally, the tight connection between BMECs minimizes intercellular
substance exchange. When the blood–brain barrier is compromised, it often results in the
infiltration of inflammatory mediators and toxic substances from the bloodstream into the
brain [187]. This can occur through the cell death of brain microvascular endothelial cells,
disruption of the tight junction structure, or a combination of both mechanisms, leading
to a cascade of brain injuries.
Figure 2. Structure of the blood–brain barrier and its intricate relationship with neuroinflammation
and ferroptosis. Serving as a crucial barrier between brain capillaries and the brain parenchyma, the
blood–brain barrier comprises brain endothelial cells, pericytes, and astrocytes, with a particular
emphasis on the tight junctions between brain endothelial cells. In instances of brain inflammation,
microglia are the primary instigators, releasing a plethora of inflammatory factors that disrupt the
internal environment of various cell types within the blood–brain barrier, leading to elevated levels
of iron and reactive oxygen species (ROS). Brain endothelial cells modulate the labile iron pool
through mechanisms such as TF/TFR, FtMt, and FPN, while astrocyte-secreted hepcidin can impact
the FPN activity of these endothelial cells. The accumulation of iron ultimately triggers ferroptosis
in the brain endothelial cells. Moreover, EMMPRIN on the cell surface activates the NF-κB pathway,
facilitating the secretion of MMP2/9, which in turn degrades tight junction proteins like occludin.
The resultant ferroptosis of endothelial cells and the breakdown of tight junction proteins are pri-
mary contributors to the compromise of the blood–brain barrier integrity. This breach allows in-
flammatory factors from the bloodstream and peripheral immune cells to infiltrate the brain paren-
Figure 2. Structure of the blood–brain barrier and its intricate relationship with neuroinflammation
and ferroptosis. Serving as a crucial barrier between brain capillaries and the brain parenchyma, the
blood–brain barrier comprises brain endothelial cells, pericytes, and astrocytes, with a particular
emphasis on the tight junctions between brain endothelial cells. In instances of brain inflammation,
microglia are the primary instigators, releasing a plethora of inflammatory factors that disrupt
the internal environment of various cell types within the blood–brain barrier, leading to elevated
levels of iron and reactive oxygen species (ROS). Brain endothelial cells modulate the labile iron
pool through mechanisms such as TF/TFR, FtMt, and FPN, while astrocyte-secreted hepcidin can
impact the FPN activity of these endothelial cells. The accumulation of iron ultimately triggers
ferroptosis in the brain endothelial cells. Moreover, EMMPRIN on the cell surface activates the
NF-
κ
B pathway, facilitating the secretion of MMP2/9, which in turn degrades tight junction proteins
like occludin. The resultant ferroptosis of endothelial cells and the breakdown of tight junction
proteins are primary contributors to the compromise of the blood–brain barrier integrity. This breach
allows inflammatory factors from the bloodstream and peripheral immune cells to infiltrate the brain
parenchyma, exacerbating neuroinflammation and eventual cell death within the brain parenchyma.
Abbreviations: Fpn, ferroportin; FtMt, ferritin mitochondrial; MMP2, matrix Metallopeptidase 2;
MMP9, matrix Metallopeptidase 9; NF-
κ
B, nuclear factor-kappaB; ROS, reactive oxygen species; TF,
transferrin; TFR, transferrin receptor. Solid arrows indicate promotion of expression or translocation.
Inhibition arrows denote suppression of expression. Dotted arrows represent chemical names.
Upward vertical arrows signify an increase in substance content. This figure was created with
BioRender.com.
Antioxidants 2024,13, 395 14 of 31
Understanding the molecular mechanisms underlying blood–brain barrier damage
and its consequences on the central nervous system is crucial due to the heavy reliance of
the central nervous system on the physiological state of the neurovascular unit (NVU) in the
brain [
188
]. Ferroptosis and neuroinflammation have emerged as significant mechanisms
implicated in the progression and outcomes of blood–brain barrier injury during disease
processes [
189
]. By comprehending these mechanisms, we can gain valuable insights
into the pathogenesis of central nervous system diseases and potentially develop targeted
therapeutic strategies.
Ferroptosis in brain microvascular endothelial cells plays a pivotal role in the compro-
mised integrity of the blood–brain barrier. This intricate process encompasses a multitude
of metabolic pathways and regulatory mechanisms, including iron metabolism, lipid
metabolism, glutathione metabolism, and the GPX4 pathway [
189
]. Disruption in any of
the components related to iron uptake, storage, or transport can elevate the labile iron pool
in BMECs, resulting in heightened permeability of the blood–brain barrier. Consequently,
BMECs transfer excess iron to neurons and glial cells in the brain via the transferrin and its
receptor (TF/TFR1) mechanism [
189
] (Figure 2). Notably, mitochondrial ferritin (FtMt) in
BMECs plays a crucial role in maintaining iron homeostasis by regulating iron storage [
190
]
(Figure 2). Research has indicated that ferroportin 1 (FPN1), an iron transporter in BMECs,
is regulated by hepcidin in astrocytes, underscoring the intricate intercellular regulation
within the blood–brain barrier [
191
]. Furthermore, the use of the iron chelating agent DFO
has been shown to reduce the labile iron pool in BMECs and protect the blood–brain bar-
rier through the HIF2
α
-Ve-Cadherin pathway [
192
]. Additionally, elevated levels of lipid
peroxidation products, such as MDA and 4HNE, have been observed in association with
blood–brain barrier disruption, concomitant with depleted GSH levels, emphasizing the
potential significance of REDOX regulation in preserving the integrity of the blood–brain
barrier [189].
In addition to BMECs, other components of the blood–brain barrier have also been
implicated in ferroptosis and its associated metabolic pathways. Notably, in the context
of Alzheimer’s disease treatment, A
β
1-40 has been found to facilitate iron overload and
lipid peroxidation in pericytes, thereby inducing blood–brain barrier impairment through
pericyte-mediated ferroptosis [
193
]. Moreover, lipid peroxidation can have an effect on the
blood–brain barrier by downregulating tight junction proteins, further compromising its
integrity [184].
In the study of different central nervous system diseases, researchers have discovered
that brain microvascular endothelial cells exhibit heightened sensitivity to ferroptosis
compared to other cell types [
194
]. In various models of brain injury resulting from cerebral
ischemia and hypoxia, Alzheimer’s disease, traumatic brain injury, and similar conditions,
ferroptosis triggered by BMECs has emerged as a crucial factor contributing to the loss of
blood–brain barrier integrity [
194
196
]. Considering the interplay between blood–brain
barrier cells and the impact of tight junction proteins affected by metabolites associated with
cell ferroptosis, it can be inferred that cell ferroptosis represents a significant underlying
cause of blood–brain barrier disruption.
Most central nervous system diseases are associated with neuroinflammatory cascades,
and the relationship between cerebral neuroinflammation and the blood–brain barrier
is bidirectional. Neuroinflammation contributes to an increase in blood–brain barrier
permeability due to various factors, while dysfunction of the blood–brain barrier leads
to an amplification of the neuroinflammatory response [
52
]. At the core of this positive
feedback loop is microglia, which can initiate inflammatory pathways following brain
damage from diverse causes, impact the structure of BMECs, and disrupt the tight junctions
of the blood–brain barrier through the release of inflammatory cytokines. Once the cell–cell
connections of the blood–brain barrier are compromised, the infiltration of peripheral
immune cells and exogenous neurotoxic substances further stimulate microglia [
52
,
197
].
This sequence of events encompasses several classic inflammatory manifestations, including
the release of TNF-
α
, IL-6, and IL-1
β
, as well as the transition of microglia from the M2
Antioxidants 2024,13, 395 15 of 31
to M1 phenotype [
198
,
199
]. In a particular study, researchers utilized lipopolysaccharide-
induced neuroinflammation to examine the microstructure of the blood–brain barrier.
Their findings revealed morphological changes in brain endothelial cells (BEC), such
as increased vesicles, plasma membrane shrinkage, mitochondrial abnormalities, and
extracellular expulsion. Additionally, notable oxidative stress characteristics were observed
in BECs [
200
]. Whether these morphological and functional alterations are linked to the
ferroptosis of BECs warrants further investigation.
In the progression of neuroinflammation, the destruction of the blood–brain bar-
rier leads to the invasion of peripheral circulating immune cells and other harmful sub-
stances, which is a significant mechanism for exacerbating brain injury (Figure 2). First,
the chemokines and cytokines released by peripheral immune cells can directly breach the
blood–brain barrier and activate the inflammatory response in the brain parenchyma [
200
].
Secondly, the infiltration of the peripheral immune cell barrier into the brain can lead to
damage of the original neurovascular unit (NVU) [
52
]. Finally, the proinflammatory pheno-
typic changes in microglia are also directly or indirectly associated with the infiltration of
peripheral cells [197].
Indeed, matrix metalloproteinases (MMPs), particularly MMP-2 and MMP-9, play a
crucial role in the mechanism of blood–brain barrier damage [
201
]. These MMPs can be
induced by extracellular matrix metalloproteinase inducer (EMMPRIN). Elevated levels of
MMP-2/9 can lead to blood–brain barrier dysfunction by degrading tight junction proteins
like ZO-1 and occludin that maintain cellular integrity [
202
]. There are various regulatory
mechanisms for MMP-2/9. Evidence suggests that lipid peroxidation in endothelial cells
can upregulate MMP-2/9, leading to the breakdown of occludin [
184
]. Additionally,
activated microglia can stimulate MMP-2/9 by increasing EMMPRIN expression through
specific signaling pathways [
189
,
202
] (Figure 2). This information highlights how both
the process of ferroptosis in cells and neuroinflammatory manifestations can contribute
to blood–brain barrier damage through MMPs. EMMPRIN and MMP-2/9 may serve as
important links connecting neuroinflammation and cellular ferroptosis in the context of
the blood–brain barrier, making them potential targets for the treatment of central nervous
system diseases.
The molecular mechanisms underlying the onset and progression of blood–brain
barrier (BBB) injury are intricate, given its crucial role in separating and connecting the
central nervous system with the peripheral blood circulation. Following the emergence
of central nervous system diseases, neuroinflammation and the occurrence of ferroptosis
in brain microvascular endothelial cells (BMECs) and other cell types contribute to the
breakdown of the BBB’s essential structure. This heightened permeability of the BBB
not only fosters a cycle of neuroinflammatory responses but also gives rise to diverse
forms of neurovascular unit (NVU) cell death in the brain. In essence, ferroptosis and
neuroinflammation can act as both instigators and outcomes of BBB injury, respectively.
Pertinent inquiries arise: (a) What specific mechanisms underpin cell ferroptosis in causing
BBB damage? (b) Are the phenotypic changes in BMECs, driven by neuroinflammation
subsequent to central nervous system diseases, linked to its ferroptosis mechanism? (c) Is
the induced NVU cell death following BBB injury directly associated with alterations in
BMECs and tight junctions (TJs)?
4.2.2. Microglia
Microglia are a distinct type of glial cell found in the central nervous system, primarily
localized around neurons and the cerebrovasculature, and their population is relatively
small compared to other glial cells. Under normal, homeostatic conditions, resting microglia
serve essential physiological functions such as promoting nerve growth and constructing
synapses [
203
]. However, when the brain experiences injury or neurodegeneration, this del-
icate balance within the central nervous system is disrupted, leading to morphological and
functional changes in microglia as they become activated [48]. The activated microglia ex-
hibit diverse characteristics, and in order to facilitate the study of their function, researchers
Antioxidants 2024,13, 395 16 of 31
have traditionally classified their activation morphology into the pro-inflammatory M1
phenotype and the anti-inflammatory M2 phenotype [
204
]. This marked functional shift
may be attributed to differing metabolic profiles in the two phenotypes [
204
]. Throughout
the progression of central nervous system diseases, certain pathogenic molecules can in-
duce the transformation of microglia into the M1 phenotype. These pathogenic molecules
include pathogen-associated molecular patterns (PAMPs) such as lipopolysaccharides,
cytokines like IFNγand GM-CSF, or DAMPs such as HMGB-1 or alpha-SYN [203].
Significantly, ferroptosis is intricately linked to phenotypic changes in microglia.
Firstly, iron metabolism in microglia plays a critical role in influencing its phenotypic
shifts. Research indicates that during the acute phase of Alzheimer’s disease or LPS-
induced neuroinflammation, the upregulation of divalent metal transporter 1 (DMT1) on the
microglial cell membrane results in iron overload and M1 polarization in microglia [
136
,
205
].
The activation of M1 phenotype microglia leads to the release of inflammatory factors such
as TNF-
α
, IL-1
β
, and IL-6, exacerbating neuroinflammation [
102
,
138
]. The classical iron
chelating agent DFO has been shown to help maintain iron homeostasis and partially
reverse M1 polarization of microglia [126].
Secondly, lipid peroxidation can induce the polarization of microglia towards the
M1 phenotype, thereby triggering microglia-mediated neuroinflammation [
206
]. Notably,
the signature molecule of ferroptosis, Acyl-CoA synthetase long-chain family member 4
(ACSL4), which regulates polyunsaturated fatty acids, has been implicated in enhancing
the release of inflammatory factors in microglia during stroke. The use of the ferroptosis
inhibitor liproxstatin-1 has demonstrated efficacy in reducing neuroinflammation resulting
from increased ACSL4 expression [
26
]. However, while ACSL4 in microglia does con-
tribute to the production of inflammatory cytokines, studies suggest that this mechanism
may be independent of ferroptosis. Data indicate that ACSL4 influences neuroinflam-
mation by affecting vestigial-like family member 4 (VGLL4), with no significant changes
in lipid peroxidation and cell-ferroptosis-related indicators observed following ACSL4
knockdown [206].
Finally, REDOX metabolism and related glutathione signaling pathways play a cru-
cial role in influencing phenotypic changes in microglia. Reduced levels of GPX4 in the
brain after injury result in the upregulation of reactive oxygen species (ROS), which in
turn enhances the pro-inflammatory response of microglia through the NF-
κ
B signaling
pathway [
126
,
135
]. Studies involving mice with GPX4 gene knockout have demonstrated
increased microglial activation and upregulation of neuroinflammatory markers [127].
Current research on the mechanism of ferroptosis on microglia phenotype has certain
limitations. This mechanism may be related to the varying levels of inducible nitric oxide
synthase (iNOS) in different cell phenotypes [
175
]. Regardless of whether it is the M1 or M2
phenotype, microglia can resist ferroptosis through the iNOS/nitric oxide (NO
) pathway.
However, the expression of iNOS in microglia of the M2 phenotype is relatively low, which
leads to the death of the M2 phenotype under conditions of iron overload [
176
]. Following
subarachnoid hemorrhage, the administration of L-NIL, an inhibitor of iNOS, can reverse
the number of M1 phenotypes in microglia and mitigate the extent of damage caused by
SAH [177].
In addition, the inflammatory effect of microglia is reflected in its role in mediating
the death of other nerve cells within the central nervous system. Microglia can disrupt the
blood–brain barrier by releasing pro-inflammatory factors that damage brain microvascular
endothelial cells and tight junction proteins [
52
]. Microglia activated prior to astrocytes
can influence astrocyte A1/A2 phenotypes through pro-inflammatory/anti-inflammatory
factors, thereby influencing astrocyte outcomes [
207
]. In mice with Parkinson’s disease,
microglia can induce ferroptosis in neuronal cells [107].
Microglia, as the core cells of central nervous system inflammation, play a crucial role
in mediating and perpetuating inflammation. They can be activated and migrate in response
to chemokines and cytokines released during inflammation, subsequently releasing pro-
inflammatory factors that contribute to the inflammatory cascade. Consequently, cell death
Antioxidants 2024,13, 395 17 of 31
and phenotypic changes within microglia themselves become significant factors in the
progression of inflammation. Several studies have demonstrated that the sensitivity of
microglia to ferroptosis, depending on their phenotypes, influences their polarization
during central nervous system inflammation, thereby exerting a comprehensive impact
on neuroinflammation. Furthermore, as cells undergoing ferroptosis release DAMPs that
promote neuroinflammation [
127
], the involvement of ferroptosis in other related cells
affected by microglia contributes to the pathological changes observed in central nervous
system diseases. However, the mechanisms through which microglia induce ferroptosis in
themselves and other nerve cells warrant further investigation.
4.2.3. Astrocytes
Astrocytes, the largest glial cell type, are ubiquitously present across mammalian
species. These cells play a pivotal role within the central nervous system, where their
primary functions span from maintaining ion homeostasis, metabolizing neurotransmitters,
and providing neurotrophic support to contributing to the integrity of the blood–brain
barrier [
53
,
208
210
]. Recent years have seen a surge in interest regarding the dynamic inter-
action between astrocytes and neurons. Beyond merely creating an optimal physiological
milieu for neuronal function, astrocytes emerge as autonomous information-processing
units that significantly influence the central nervous system’s overarching functional-
ity [211].
Astrocyte lesions play a critical role in a wide array of central nervous system dis-
eases, including traumatic brain injury, neurodegenerative disorders, and cerebrovascular
diseases. The pathophysiological changes observed in astrocytes during these conditions
typically include morphological atrophy, various modes of cell death, and the proliferation
of reactive astrocytes [
212
]. Among these, reactive astrocytosis has garnered significant
attention. Initially, much like microglia, reactive astrocytes were categorized into neurotoxic
(A1) and neuroprotective (A2) types, reflecting their dual roles in neural environments [
204
].
However, emerging perspectives suggest that astrocytes’ response to disease and trauma
is predominantly adaptive, thus primarily serving a neuroprotective function [
212
]. This
shift in understanding underscores the importance of focusing research on the functional
phenotypes of reactive glial cells rather than a binary classification, offering new insights
into astrocytes’ involvement in neuroinflammation.
Current research indicates that astrocytes undergo phenotypic transformations that
play a crucial role in modulating neuroinflammation, leading to the production of various
inflammatory cytokines and cytotoxic reactive oxygen species (ROS) [
49
]. Interestingly,
astrocytes can generate antioxidant enzymes under pathological conditions to mitigate ox-
idative stress [
49
]. During inflammation, astrocytes’ regulation of mitochondrial functions,
ROS, and antioxidant systems is pivotal in controlling their susceptibility to ferropto-
sis [
213
]. Building on this understanding and the intricate material exchanges between
astrocytes and neurons, we delve into the connection between astrocyte-driven neuroin-
flammation and ferroptosis, highlighting the complex interplay that influences disease
progression and potential therapeutic targets.
During the neuroinflammation of the telencephalon, astrocytes become activated,
identifiable by specific biomarkers such as GFAP, S100
β
, and CD81 [
49
,
214
]. The function-
ality of astrocytes is intricately linked to that of microglia, with various proinflammatory
and anti-inflammatory cytokines and chemokines serving as crucial mediators of their
interaction. Both cell types are capable of influencing the other’s phenotype through the
release of factors like TNF-
α
, which promotes inflammation, or TGF-
β
, which dampens it,
thereby either exacerbating or mitigating the inflammatory response [
49
,
204
]. Furthermore,
activated astrocytes can produce matrix metalloproteinases (MMPs), compromising the
blood–brain barrier and facilitating the influx of peripheral inflammatory cells into the
central nervous system, thereby intensifying the inflammatory milieu within the brain [
49
].
The release of pro-inflammatory and anti-inflammatory mediators during inflamma-
tion, coupled with oxidative stress-related enzymes, triggers the activation of intracellular
Antioxidants 2024,13, 395 18 of 31
receptors and signaling pathways. Notably, NADPH oxidase 4 (NOX4) in astrocytes
catalyzes the production of reactive oxygen species (ROS), further promoting neuroin-
flammation and mitochondrial damage [
215
]. This ROS-induced activity leads to lipid
peroxidation within cells, elevating levels of 4HNE and MDA and initiating ferroptosis in
astrocytes [
213
,
216
]. The activation of ROS also correlates with increased concentrations
of IL-1
β
, IL-6, TNF-
α
, and other inflammatory cytokines, suggesting a temporal over-
lap between the induction of astrocyte inflammation and ferroptosis. The upregulation
of glutathione peroxidase 4 in astrocytes acts as a defense mechanism against oxidative
stress-induced ferroptosis [49].
Moreover, the proinflammatory phenotype of astrocytes appears to be closely associ-
ated with the NF-
κ
B signaling pathway. The interplay between Nrf2 and NF-
κ
B pathways
is central to the molecular mechanisms underlying neurological diseases. Nrf2 can miti-
gate the expression of reactive astrocytes and reduce neuroinflammation and ferroptosis
by inhibiting NF-
κ
B [
217
,
218
]. The regulator of G protein signaling 5 (RGS5) can induce
neuroinflammation via the tumor necrosis factor receptor (TNFR) signaling pathway, con-
tributing to neurodegeneration in conditions like Parkinson’s disease, with TLR4 mediating
downstream factors of this pathway and regulating NF-
κ
B activation through nuclear
translocation [
131
,
219
]. Experiments treating angiotensin II (Ang II)-stimulated astrocytes
with the ferroptosis-specific inhibitor ferrostatin-1 have shown concurrent alterations in
astrocyte markers, inflammatory factors, and ferroptosis-related proteins [
58
]. Therefore,
there may be a common regulatory approach between astrocyte-induced neuroinflamma-
tion and cellular ferroptosis.
Furthermore, the process of inflammatory induction in astrocytes can trigger ferropto-
sis in neurons by altering the internal milieu of the central nervous system. Chemokines
derived from astrocyte activation, such as CXCL10/CXCR3, and the inflammatory cytokine
IL-6 can alter the expression of neuronal proteins, including Fpn, leading to an increase in
intracellular iron levels and ultimately resulting in neuronal ferroptosis [
29
,
220
]. However,
recent studies have indicated that the regulation of ferroptosis in neurons by astrocytes is
complex. Astrocytes may leverage the protective potential of Nrf2 to counter ferroptosis in
neurons by releasing exosomes [221].
4.2.4. Neutrophils and NETs
Neutrophils, originating from the bone marrow, are widely distributed in the blood-
stream and play a crucial role in acute inflammation, being the most prevalent cells during
its initial stages. Evidence suggests that the Neutrophil to Lymphocyte Ratio (NLR) is
associated with prognosis in certain central nervous system diseases [
222
224
]. How-
ever, the physiological blood–brain barrier presents a challenge for neutrophil entry into
the telencephalon. Consequently, the contribution of peripheral cells to central nervous
system diseases is often underestimated. Nevertheless, various pathological conditions
can compromise the blood–brain barrier, eventually allowing neutrophil infiltration. The
interplay between neutrophils and blood–brain barrier integrity is an unavoidable subject
in central nervous system diseases. Both systemic circulation and inflammatory factors in
the brain milieu can contribute to blood–brain barrier disruption [
225
]. Under pathologi-
cal conditions, chemokines and cytokines can drive neutrophil aggregation towards the
blood–brain barrier, further exacerbating its integrity [
226
]. Once the blood–brain barrier is
breached, neutrophils can infiltrate the brain, triggering neuroinflammation and damage to
the nervous system.
Recent advances have reshaped the understanding of neutrophils, highlighting their
diverse phenotypes and functions, contrary to the prior notion of their homogeneity [
227
].
Different neutrophil phenotypes play distinct roles in brain neuroinflammation [
228
].
Studies on neutrophil aging indicate that prolonged neutrophil survival amplifies their
pro-inflammatory impact on the brain [
229
]. Over the years, research on Neutrophil
Extracellular Traps (NETs) in neurological disorders has surged. NETs represent the
unique response of mature neutrophils to trauma and pathogens, forming a meshwork
Antioxidants 2024,13, 395 19 of 31
structure for self-destruction [
230
]. Co-culturing of NETs with nerve cells has demonstrated
their ability to provoke nerve cell inflammation. Additionally, ferroptosis in neutrophils
contributes to NET formation [
231
]. The inhibition of NET formation by ferroptosis inhibitor
ferrostatin-1 underscores the intimate link between ferroptosis and NET generation in
neutrophils [
231
]. Nonetheless, there is a lack of
in vivo
evidence of neutrophil ferroptosis
post-infiltration into brain tissue through the compromised blood–brain barrier, triggering
NET release and exacerbating neuroinflammation. Studies have shown that NETs modulate
ferroptosis in alveolar epithelium and breast cancer cell death through NF-
κ
B-related
pathways [
232
,
233
]. The intricate interplay among reactive oxygen species (ROS), NF-
κ
B,
and other inflammatory factors in inducing neuroinflammation and ferroptosis presents a
compelling avenue for exploring the processes and molecular mechanisms of NET-driven
ferroptosis and neuroinflammation in the brain.
Furthermore, both inflammatory cells and ferroptotic cells can release hypermigratory
histone box 1 (HMGB1), a protein that plays a crucial role in neutrophil recruitment [
234
].
Based on the aforementioned discoveries, we propose several hypotheses: In certain central
nervous system disorders, such as degenerative diseases and acute-phase conditions like
trauma, there may exist a detrimental cycle involving intracerebral nerve cells, peripheral
neutrophils, and intracerebral nerve cells. The release of DAMPs and inflammatory factors
during neuroinflammation of brain cells can lead to neutrophil recruitment and disruption
of the blood–brain barrier. Consequently, neutrophil infiltration into brain tissue occurs,
triggering their ferroptosis. Following ferroptosis, neutrophils release NETs, which can
further exacerbate the inflammatory response and neuronal death through interactions
with neurons.
4.2.5. T Cells
Lymphocytes are pivotal in orchestrating immune responses within the body. Based
on their morphology and function, they can be classified into B lymphocytes, T lympho-
cytes, and natural killer (NK) cells. In certain neurological conditions that incite chronic
neuroinflammation, both B and T cells have been observed to infiltrate the cerebrospinal
fluid, meninges, blood–brain barrier, and brain parenchyma [
51
,
235
237
]. Nevertheless,
extensive research indicates that T lymphocytes predominantly contribute to the pathogen-
esis of central nervous system diseases. B cells, on the other hand, play a crucial role in
maintaining the microenvironment and activating and sustaining the pro-inflammatory
function of T cells [235,236].
It is important to note that chemokines play a crucial role in the infiltration and
proliferation of T cells and B cells within the central nervous system. Chemokines can
be classified into four different subgroups: CXC, CC, CX3C, and XC [
238
]. Lymphocytes
express various chemokine receptors, which enable them to access different sites via the
chemokine ligand/receptor axis when chronic neuroinflammation occurs [239].
T lymphocytes are categorized into cytotoxic T cells (Tc cells), helper T cells (Th
cells), and regulatory T cells (Treg cells) based on their functions. These T lymphocytes
play a significant role in immune-related and chronic infectious neurological diseases by
releasing cytokines. Helper T cells, specifically Th1 and Th17 cells, are strongly associated
with neuroinflammation in the central nervous system. Research has demonstrated that
CNS astrocytes respond to cytokines released during Th1 and Th17 cell activation [
240
].
Th1 cells can secrete IFN-
γ
, which promotes the infiltration of T cells into the spinal cord.
Meanwhile, astrocytes partially inhibit T cell infiltration through the VCAM-1 receptor [
240
].
In the brain, the invasion of Th cells involves antigen-presenting cells (APCs) such as
boundary-associated macrophages (BAMs), classical dendritic cells (cDCs), and B cells [
237
].
Activation of Th1 and Th17 cells plays a critical role in promoting neuroinflammation by T
cells. Jinyuan et al. discovered that in multiple sclerosis, neurons undergoing ferroptosis
stimulate the activation of T cell receptor signaling pathways through the secretion of
various related molecules, leading to the activation of Th1 and Th17 cells and the release of
cytokines like IFN-γand IL-17 [241].
Antioxidants 2024,13, 395 20 of 31
CD8+ T cells are crucial cytotoxic T cells that have destructive effects on both brain
parenchyma and blood vessels. In chronic neuroinflammation, CD8+ T cells can infiltrate
and destroy blood vessels. In pathological models of cerebral malaria and Susac syndrome,
CD8+ T cells can damage cerebral vascular endothelial cells and decrease tight junction
protein expression via adhesion molecules such as VCAM-1 [
242
]. The breakdown of the
blood–brain barrier by CD8+ T cells ultimately leads to their infiltration into the brain
parenchyma [
51
]. Activated CD8+ T cells play a vital role in inducing ferroptosis in neurons,
which is confirmed by upregulation of related proteins such as TFR1 and ACSL4 in ECM
brain models [
243
]. Th cell activation by neuronal ferroptosis may locally exacerbate the
“vicious circle” of neuroinflammation [
51
]. Additionally, the infiltration of CD8+ T cells in
tau transgenic mice may activate microglia and astrocytes on the surface, leading to further
aggravation of neuroinflammation in the brain [
244
]. In contrast, Treg cells often display
neuroprotective effects, and their proliferation can be promoted by IL-2 secreted by helper
T cells [245]. Depleting Treg cells in cases of brain injury is not favorable for recovery.
Peripheral immune cells represent a promising area of research in the context of central
nervous system inflammation. While previous studies of CNS diseases have primarily
focused on the interaction between neurons and glial cells in situ, the significance of periph-
eral immune cells in the blood–brain barrier and brain parenchyma cannot be overlooked.
Particularly in certain chronic neuroinflammation disease models, there is evidence sug-
gesting that neuronal ferroptosis may both result from and contribute to the aggregation
and activation of T cells. The infiltration of immune cells exacerbates the disruption of
the nervous system’s physiological equilibrium and intensifies neuroinflammation. The
quantity and function of lymphocytes in numerous central nervous system disease models
warrant further investigation.
5. Conclusions
This review comprehensively summarizes the molecular mechanisms and cellular re-
sponses linked to ferroptosis and neuroinflammation in central neurological disorders. Our
literature search revealed a potential interplay between the metabolites of ferroptosis and
neuroinflammation, with each potentially acting as the initiator of the other process. While
the detailed pathogenesis varies, various central neurological disorders involve immune
cell activation in the brain, disruption of the blood–brain barrier, intracranial iron overload,
and peroxidation. These changes underscore the close correlation between ferroptosis and
neuroinflammation in the spatial and temporal distribution of each disease. Furthermore,
we focused on elucidating the specific mechanisms involving SIRT1, Nrf2, NF-
κ
B, Cox-2,
and iNOS/NO
molecules upstream and downstream of each other in central neurolog-
ical disorders, highlighting their connections with ferroptosis and neuroinflammation.
However, the relationships between these molecules are intricate, with many intermediate
processes still not fully understood. Moreover, the varying content of these regulatory
molecules in different nervous system cells significantly influences the susceptibility of
different cell phenotypes to ferroptosis and neuroinflammation. Future research on these
mechanisms could delve deeper into the following aspects: a. Intermediate processes
linking these associated molecules. b. Variances in the impact of different molecules over
time and across various cell phenotypes. c. The influence of peripheral immune cells on
the central nervous system environment.
6. Futures Perspectives
Given the intricate structure of the human brain and the diverse functions of nerve
cells, the treatment options for many central neurological disorders are currently limited.
Consequently, researchers are concentrating on exploring new molecular mechanisms
underlying these diseases. Ferroptosis, identified as a specific form of cell death, has
been implicated in various neurological disorders. By elucidating the interactive signaling
pathways and critical regulatory molecules connecting ferroptosis and neuroinflammation,
we have broadened the spectrum of potential therapeutic targets aimed at mitigating
Antioxidants 2024,13, 395 21 of 31
neurological damage in central nervous system disorders. For instance, resveratrol, which
modulates SIRT1, shows promise as a viable treatment for a range of neuroinflammatory
conditions [
93
]. Undoubtedly, further endeavors are imperative to translate anti-ferroptosis
or anti-inflammatory drugs into clinical practice. Apart from conducting phased animal
studies and clinical trials, enhancing the integration between the two pathways is equally
pivotal.
Author Contributions: Conceptualization, Y.X., Q.L. and C.L.; Writing—Original Draft Preparation,
Y.X. and C.L.; Review and Editing, B.J., J.L. and Q.L.; Funding Acquisition, C.L. All authors have read
and agreed to the published version of the manuscript.
Funding: This research was supported by the National Natural Science Foundation of China
(82271409, 81971163), Major project of National Natural Science Foundation of China (82293650,
82293651), Open project of Hebei Key Laboratory of Forensic Medicine (JYFY-23KF002), the outstand-
ing young backbone teacher of the “Blue Project” in Jiangsu Province (SR13450121), and a Project
Funded by the Priority Academic Program Development of Jiangsu Higher Education Institutions
(PAPD).
Conflicts of Interest: The authors declare no conflicts of interest.
References
1.
Sauve, F.; Nampoothiri, S.; Clarke, S.A.; Fernandois, D.; Ferreira Coelho, C.F.; Dewisme, J.; Mills, E.G.; Ternier, G.; Cotellessa, L.;
Iglesias-Garcia, C.; et al. Long-COVID cognitive impairments and reproductive hormone deficits in men may stem from GnRH
neuronal death. EBioMedicine 2023,96, 104784. [CrossRef] [PubMed]
2.
Choi, B.Y.; Hong, D.K.; Kang, B.S.; Lee, S.H.; Choi, S.; Kim, H.J.; Lee, S.M.; Suh, S.W. Engineered Mesenchymal Stem Cells
Over-Expressing BDNF Protect the Brain from Traumatic Brain Injury-Induced Neuronal Death, Neurological Deficits, and
Cognitive Impairments. Pharmaceuticals 2023,16, 436. [CrossRef] [PubMed]
3.
Pang, K.; Jiang, R.; Zhang, W.; Yang, Z.; Li, L.L.; Shimozawa, M.; Tambaro, S.; Mayer, J.; Zhang, B.; Li, M.; et al. An App knock-in
rat model for Alzheimer’s disease exhibiting Abeta and tau pathologies, neuronal death and cognitive impairments. Cell Res.
2022,32, 157–175. [CrossRef] [PubMed]
4.
Dixon, S.J.; Lemberg, K.M.; Lamprecht, M.R.; Skouta, R.; Zaitsev, E.M.; Gleason, C.E.; Patel, D.N.; Bauer, A.J.; Cantley, A.M.; Yang,
W.S.; et al. Ferroptosis: An iron-dependent form of nonapoptotic cell death. Cell 2012,149, 1060–1072. [CrossRef] [PubMed]
5.
Chen, D.; Fan, Z.; Rauh, M.; Buchfelder, M.; Eyupoglu, I.Y.; Savaskan, N. ATF4 promotes angiogenesis and neuronal cell death
and confers ferroptosis in a xCT-dependent manner. Oncogene 2017,36, 5593–5608. [CrossRef] [PubMed]
6.
Zhang, Y.; Shaabani, S.; Vowinkel, K.; Trombetta-Lima, M.; Sabogal-Guaqueta, A.M.; Chen, T.; Hoekstra, J.; Lembeck, J.; Schmidt,
M.; Decher, N.; et al. Novel SK channel positive modulators prevent ferroptosis and excitotoxicity in neuronal cells. Biomed.
Pharmacother. 2024,171, 116163. [CrossRef] [PubMed]
7.
Cagnin, A.; Gerhard, A.; Banati, R.B.
In vivo
imaging of neuroinflammation. Eur. Neuropsychopharmacol. 2002,12, 581–586.
[CrossRef] [PubMed]
8.
Adrian, M.; Weber, M.; Tsai, M.C.; Glock, C.; Kahn, O.I.; Phu, L.; Cheung, T.K.; Meilandt, W.J.; Rose, C.M.; Hoogenraad, C.C.
Polarized microtubule remodeling transforms the morphology of reactive microglia and drives cytokine release. Nat. Commun.
2023,14, 6322. [CrossRef]
9.
Jiang, W.; Wang, Y.; Sun, W.; Zhang, M. Morin Suppresses Astrocyte Activation and Regulates Cytokine Release in Bone Cancer
Pain Rat Models. Phytother. Res. 2017,31, 1298–1304. [CrossRef] [PubMed]
10.
Fang, D.; Guo, S.; Wei, B.; Liu, W.; Li, G.; Li, X.; Liu, J.; Jin, L.; Duan, C. Nrf-2 modulates excitability of hippocampal neurons
by regulating ferroptosis and neuroinflammation after subarachnoid hemorrhage in rats. Brain Res. Bull. 2024,207, 110877.
[CrossRef] [PubMed]
11.
Nemeth, E.; Rivera, S.; Gabayan, V.; Keller, C.; Taudorf, S.; Pedersen, B.K.; Ganz, T. IL-6 mediates hypoferremia of inflammation
by inducing the synthesis of the iron regulatory hormone hepcidin. J. Clin. Investig. 2004,113, 1271–1276. [CrossRef] [PubMed]
12.
Dong, B.; Jiang, Y.; Shi, B.; Zhang, Z.; Zhang, Z. Selenomethionine alleviates decabromodiphenyl ether-induced oxidative stress
and ferroptosis via the NRF2/GPX4 pathway in the chicken brain. J. Hazard. Mater. 2024,465, 133307. [CrossRef] [PubMed]
13.
Mumbauer, S.; Pascual, J.; Kolotuev, I.; Hamaratoglu, F. Ferritin heavy chain protects the developing wing from reactive oxygen
species and ferroptosis. PLoS Genet. 2019,15, e1008396. [CrossRef] [PubMed]
14.
Luo, C.; Xu, W.; Tang, X.; Liu, X.; Cheng, Y.; Wu, Y.; Xie, Z.; Wu, X.; He, X.; Wang, Q.; et al. Canonical Wnt signaling works
downstream of iron overload to prevent ferroptosis from damaging osteoblast differentiation. Free Radic. Biol. Med. 2022,188,
337–350. [CrossRef] [PubMed]
15.
Hu, W.; Zhou, C.; Jing, Q.; Li, Y.; Yang, J.; Yang, C.; Wang, L.; Hu, J.; Li, H.; Wang, H.; et al. FTH promotes the proliferation and
renders the HCC cells specifically resist to ferroptosis by maintaining iron homeostasis. Cancer Cell Int. 2021,21, 709. [CrossRef]
[PubMed]
Antioxidants 2024,13, 395 22 of 31
16.
Bao, X.; Luo, X.; Bai, X.; Lv, Y.; Weng, X.; Zhang, S.; Leng, Y.; Huang, J.; Dai, X.; Wang, Y.; et al. Cigarette tar mediates macrophage
ferroptosis in atherosclerosis through the hepcidin/FPN/SLC7A11 signaling pathway. Free Radic. Biol. Med. 2023,201, 76–88.
[CrossRef] [PubMed]
17.
Yang, S.; Zhang, T.; Ge, Y.; Cheng, Y.; Yin, L.; Pu, Y.; Chen, Z.; Liang, G. Ferritinophagy Mediated by Oxidative Stress-Driven
Mitochondrial Damage Is Involved in the Polystyrene Nanoparticles-Induced Ferroptosis of Lung Injury. ACS Nano 2023,17,
24988–25004. [CrossRef] [PubMed]
18.
Tuo, Q.Z.; Liu, Y.; Xiang, Z.; Yan, H.F.; Zou, T.; Shu, Y.; Ding, X.L.; Zou, J.J.; Xu, S.; Tang, F.; et al. Thrombin induces ACSL4-
dependent ferroptosis during cerebral ischemia/reperfusion. Signal Transduct. Target. Ther. 2022,7, 59. [CrossRef]
19.
Wu, C.; Du, M.; Yu, R.; Cheng, Y.; Wu, B.; Fu, J.; Tan, W.; Zhou, Q.; Balawi, E.; Liao, Z.B. A novel mechanism linking ferroptosis
and endoplasmic reticulum stress via the circPtpn14/miR-351-5p/5-LOX signaling in melatonin-mediated treatment of traumatic
brain injury. Free Radic. Biol. Med. 2022,178, 271–294. [CrossRef] [PubMed]
20.
Zhou, Q.; Meng, Y.; Li, D.; Yao, L.; Le, J.; Liu, Y.; Sun, Y.; Zeng, F.; Chen, X.; Deng, G. Ferroptosis in cancer: From molecular
mechanisms to therapeutic strategies. Signal Transduct. Target. Ther. 2024,9, 55. [CrossRef] [PubMed]
21.
Doll, S.; Freitas, F.P.; Shah, R.; Aldrovandi, M.; da Silva, M.C.; Ingold, I.; Goya Grocin, A.; Xavier da Silva, T.N.; Panzilius, E.;
Scheel, C.H.; et al. FSP1 is a glutathione-independent ferroptosis suppressor. Nature 2019,575, 693–698. [CrossRef]
22.
Mao, C.; Liu, X.; Zhang, Y.; Lei, G.; Yan, Y.; Lee, H.; Koppula, P.; Wu, S.; Zhuang, L.; Fang, B.; et al. DHODH-mediated ferroptosis
defence is a targetable vulnerability in cancer. Nature 2021,593, 586–590. [CrossRef] [PubMed]
23.
Hu, Q.; Wei, W.; Wu, D.; Huang, F.; Li, M.; Li, W.; Yin, J.; Peng, Y.; Lu, Y.; Zhao, Q.; et al. Blockade of GCH1/BH4 Axis Activates
Ferritinophagy to Mitigate the Resistance of Colorectal Cancer to Erastin-Induced Ferroptosis. Front. Cell Dev. Biol. 2022,10,
810327. [CrossRef] [PubMed]
24. Jacquemyn, J.; Ralhan, I.; Ioannou, M.S. Driving factors of neuronal ferroptosis. Trends Cell Biol. 2024. [CrossRef] [PubMed]
25. Selim, M.H.; Ratan, R.R. The role of iron neurotoxicity in ischemic stroke. Ageing Res. Rev. 2004,3, 345–353. [CrossRef]
26.
Cui, Y.; Zhang, Y.; Zhao, X.; Shao, L.; Liu, G.; Sun, C.; Xu, R.; Zhang, Z. ACSL4 exacerbates ischemic stroke by promoting
ferroptosis-induced brain injury and neuroinflammation. Brain Behav. Immun. 2021,93, 312–321. [CrossRef] [PubMed]
27.
Ashraf, A.; Jeandriens, J.; Parkes, H.G.; So, P.W. Iron dyshomeostasis, lipid peroxidation and perturbed expression of cys-
tine/glutamate antiporter in Alzheimer’s disease: Evidence of ferroptosis. Redox Biol. 2020,32, 101494. [CrossRef] [PubMed]
28.
Ding, X.-S.; Gao, L.; Han, Z.; Eleuteri, S.; Shi, W.; Shen, Y.; Song, Z.-Y.; Su, M.; Yang, Q.; Qu, Y.; et al. Ferroptosis in Parkinson’s
disease: Molecular mechanisms and therapeutic potential. Ageing Res. Rev. 2023,91, 102077. [CrossRef] [PubMed]
29.
Liang, P.; Zhang, X.; Zhang, Y.; Wu, Y.; Song, Y.; Wang, X.; Chen, T.; Liu, W.; Peng, B.; Yin, J.; et al. Neurotoxic A1 astrocytes
promote neuronal ferroptosis via CXCL10/CXCR3 axis in epilepsy. Free Radic. Biol. Med. 2023,195, 329–342. [CrossRef] [PubMed]
30.
Xie, B.S.; Wang, Y.Q.; Lin, Y.; Mao, Q.; Feng, J.F.; Gao, G.Y.; Jiang, J.Y. Inhibition of ferroptosis attenuates tissue damage and
improves long-term outcomes after traumatic brain injury in mice. CNS Neurosci. Ther. 2019,25, 465–475. [CrossRef] [PubMed]
31.
Schroder, N.; Figueiredo, L.S.; de Lima, M.N. Role of brain iron accumulation in cognitive dysfunction: Evidence from animal
models and human studies. J. Alzheimers Dis. 2013,34, 797–812. [CrossRef] [PubMed]
32.
Fredriksson, A.; Schröder, N.; Eriksson, P.; Izquierdo, I.; Archer, T. Neonatal iron exposure induces neurobehavioural dysfunctions
in adult mice. Toxicol. Appl. Pharm. 1999,159, 25–30. [CrossRef]
33.
Mei, H.; Wu, D.; Yong, Z.; Cao, Y.; Chang, Y.; Liang, J.; Jiang, X.; Xu, H.; Yang, J.; Shi, X.; et al. PM(2.5) exposure exacerbates
seizure symptoms and cognitive dysfunction by disrupting iron metabolism and the Nrf2-mediated ferroptosis pathway. Sci.
Total Environ. 2024,910, 168578. [CrossRef]
34.
Xie, R.; Zhao, W.; Lowe, S.; Bentley, R.; Hu, G.; Mei, H.; Jiang, X.; Sun, C.; Wu, Y.; Liu, Y. Quercetin alleviates kainic acid-induced
seizure by inhibiting the Nrf2-mediated ferroptosis pathway. Free Radic. Biol. Med. 2022,191, 212–226. [CrossRef] [PubMed]
35.
Du, L.; Wu, Y.; Jia, Q.; Li, J.; Li, Y.; Ma, H.; Fan, Z.; Guo, X.; Li, L.; Peng, Y.; et al. Autophagy Suppresses Ferroptosis by Degrading
TFR1 to Alleviate Cognitive Dysfunction in Mice with SAE. Cell Mol. Neurobiol. 2023,43, 3605–3622. [CrossRef] [PubMed]
36.
Fang, J.; Yuan, Q.; Du, Z.; Zhang, Q.; Yang, L.; Wang, M.; Yang, W.; Yuan, C.; Yu, J.; Wu, G.; et al. Overexpression of GPX4
attenuates cognitive dysfunction through inhibiting hippocampus ferroptosis and neuroinflammation after traumatic brain injury.
Free Radic. Biol. Med. 2023,204, 68–81. [CrossRef]
37.
Weiland, A.; Wang, Y.; Wu, W.; Lan, X.; Han, X.; Li, Q.; Wang, J. Ferroptosis and Its Role in Diverse Brain Diseases. Mol. Neurobiol.
2019,56, 4880–4893. [CrossRef] [PubMed]
38.
Han, T.; Xu, Y.; Sun, L.; Hashimoto, M.; Wei, J. Microglial response to aging and neuroinflammation in the development of
neurodegenerative diseases. Neural Regen. Res. 2024,19, 1241–1248. [CrossRef] [PubMed]
39.
Liang, Y.; Aditi; Onyoni, F.; Wang, H.; Gonzales, C.; Sunyakumthorn, P.; Wu, P.; Samir, P.; Soong, L. Brain transcriptomics reveal
the activation of neuroinflammation pathways during acute Orientia tsutsugamushi infection in mice. Front. Immunol. 2023,14,
1194881. [CrossRef] [PubMed]
40.
Merchak, A.R.; Cahill, H.J.; Brown, L.C.; Brown, R.M.; Rivet-Noor, C.; Beiter, R.M.; Slogar, E.R.; Olgun, D.G.; Gaultier, A. The activ-
ity of the aryl hydrocarbon receptor in T cells tunes the gut microenvironment to sustain autoimmunity and neuroinflammation.
PLoS Biol. 2023,21, e3002000. [CrossRef] [PubMed]
41.
Qi, X.H.; Chen, P.; Wang, Y.J.; Zhou, Z.P.; Liu, X.C.; Fang, H.; Wang, C.W.; Liu, J.; Liu, R.Y.; Liu, H.K.; et al. Increased
cysteinyl-tRNA synthetase drives neuroinflammation in Alzheimer’s disease. Transl. Neurodegener. 2024,13, 3. [CrossRef]
[PubMed]
Antioxidants 2024,13, 395 23 of 31
42.
Hellani, F.; Leleu, I.; Saidi, N.; Martin, N.; Lecoeur, C.; Werkmeister, E.; Koffi, D.; Trottein, F.; Yapo-Ette, H.; Das, B.; et al. Role of
astrocyte senescence regulated by the non- canonical autophagy in the neuroinflammation associated to cerebral malaria. Brain
Behav. Immun. 2023,117, 20–35. [CrossRef] [PubMed]
43.
Qi, R.; Wang, X. Inhibition of miR-429 improves neurological recovery of traumatic brain injury mice and attenuates microglial
neuroinflammation. Int. Immunopharmacol. 2020,79, 106091. [CrossRef] [PubMed]
44.
Liu, Y.; Yang, W.; Xue, J.; Chen, J.; Liu, S.; Zhang, S.; Zhang, X.; Gu, X.; Dong, Y.; Qiu, P. Neuroinflammation: The central enabler
of postoperative cognitive dysfunction. Biomed. Pharmacother. 2023,167, 115582. [CrossRef] [PubMed]
45.
Li, W.; Wu, J.; Zeng, Y.; Zheng, W. Neuroinflammation in epileptogenesis: From pathophysiology to therapeutic strategies. Front.
Immunol. 2023,14, 1269241. [CrossRef] [PubMed]
46.
Liu, P.Y.; Li, H.Q.; Dong, M.Q.; Gu, X.Y.; Xu, S.Y.; Xia, S.N.; Bao, X.Y.; Xu, Y.; Cao, X. Infiltrating myeloid cell-derived properdin
markedly promotes microglia-mediated neuroinflammation after ischemic stroke. J. Neuroinflamm. 2023,20, 260. [CrossRef]
[PubMed]
47.
Rahimian, R.; Beland, L.C.; Sato, S.; Kriz, J. Microglia-derived galectin-3 in neuroinflammation; a bittersweet ligand? Med. Res.
Rev. 2021,41, 2582–2589. [CrossRef] [PubMed]
48. Long, H.Z.; Zhou, Z.W.; Cheng, Y.; Luo, H.Y.; Li, F.J.; Xu, S.G.; Gao, L.C. The Role of Microglia in Alzheimer’s Disease From the
Perspective of Immune Inflammation and Iron Metabolism. Front. Aging Neurosci. 2022,14, 888989. [CrossRef] [PubMed]
49.
He, L.; Zhang, R.; Yang, M.; Lu, M. The role of astrocyte in neuroinflammation in traumatic brain injury. Biochim. Biophys. Acta
Mol. Basis Dis. 2024,1870, 166992. [CrossRef] [PubMed]
50.
Skok, M. Mesenchymal stem cells as a potential therapeutic tool to cure cognitive impairment caused by neuroinflammation.
World J. Stem Cells 2021,13, 1072–1083. [CrossRef] [PubMed]
51.
Gonzalez-Fierro, C.; Fonte, C.; Dufourd, E.; Cazaentre, V.; Aydin, S.; Engelhardt, B.; Caspi, R.R.; Xu, B.; Martin-Blondel, G.; Spicer,
J.A.; et al. Effects of a Small-Molecule Perforin Inhibitor in a Mouse Model of CD8 T Cell-Mediated Neuroinflammation. Neurol.
Neuroimmunol. Neuroinflamm. 2023,10, e200117. [CrossRef] [PubMed]
52.
Puech, C.; Badran, M.; Runion, A.R.; Barrow, M.B.; Cataldo, K.; Gozal, D. Cognitive Impairments, Neuroinflammation and
Blood-Brain Barrier Permeability in Mice Exposed to Chronic Sleep Fragmentation during the Daylight Period. Int. J. Mol. Sci.
2023,24, 9880. [CrossRef] [PubMed]
53.
Gharbi, T.; Zhang, Z.; Yang, G.Y. The Function of Astrocyte Mediated Extracellular Vesicles in Central Nervous System Diseases.
Front. Cell Dev. Biol. 2020,8, 568889. [CrossRef] [PubMed]
54.
Wang, C.S.; Kavalali, E.T.; Monteggia, L.M. BDNF signaling in context: From synaptic regulation to psychiatric disorders. Cell
2022,185, 62–76. [CrossRef] [PubMed]
55.
Li, Y.; Pan, K.; Chen, L.; Ning, J.L.; Li, X.; Yang, T.; Terrando, N.; Gu, J.; Tao, G. Deferoxamine regulates neuroinflammation and
iron homeostasis in a mouse model of postoperative cognitive dysfunction. J. Neuroinflamm. 2016,13, 268. [CrossRef] [PubMed]
56.
Jung, H.Y.; Kwon, H.J.; Kim, W.; Hwang, I.K.; Choi, G.M.; Chang, I.B.; Kim, D.W.; Moon, S.M. Tat-Endophilin A1 Fusion
Protein Protects Neurons from Ischemic Damage in the Gerbil Hippocampus: A Possible Mechanism of Lipid Peroxidation and
Neuroinflammation Mitigation as Well as Synaptic Plasticity. Cells 2021,10, 357. [CrossRef] [PubMed]
57.
He, K.; Nie, L.; Ali, T.; Liu, Z.; Li, W.; Gao, R.; Zhang, Z.; Liu, J.; Dai, Z.; Xie, Y.; et al. Adiponectin deficiency accelerates brain
aging via mitochondria-associated neuroinflammation. Immun. Ageing 2023,20, 15. [CrossRef] [PubMed]
58.
Li, S.; Zhou, C.; Zhu, Y.; Chao, Z.; Sheng, Z.; Zhang, Y.; Zhao, Y. Ferrostatin-1 alleviates angiotensin II (Ang II)- induced
inflammation and ferroptosis in astrocytes. Int. Immunopharmacol. 2021,90, 107179. [CrossRef] [PubMed]
59.
Li, M.; Li, S.C.; Dou, B.K.; Zou, Y.X.; Han, H.Z.; Liu, D.X.; Ke, Z.J.; Wang, Z.F. Cycloastragenol upregulates SIRT1 expression,
attenuates apoptosis and suppresses neuroinflammation after brain ischemia. Acta Pharmacol. Sin. 2020,41, 1025–1032. [CrossRef]
[PubMed]
60.
Priyanka, S.H.; Syam Das, S.; Thushara, A.J.; Rauf, A.A.; Indira, M. All Trans Retinoic Acid Attenuates Markers of Neuroinflam-
mation in Rat Brain by Modulation of SIRT1 and NFkappaB. Neurochem. Res. 2018,43, 1791–1801. [CrossRef]
61.
Song, H.; Ding, Z.; Chen, J.; Chen, T.; Wang, T.; Huang, J. The AMPK-SIRT1-FoxO1-NF-kappaB signaling pathway participates in
hesperetin-mediated neuroprotective effects against traumatic brain injury via the NLRP3 inflammasome. Immunopharmacol.
Immunotoxicol. 2022,44, 970–983. [CrossRef] [PubMed]
62.
Xu, J.; Zhao, L.; Zhang, X.; Ying, K.; Zhou, R.; Cai, W.; Wu, X.; Jiang, H.; Xu, Q.; Miao, D.; et al. Salidroside ameliorates
acetaminophen-induced acute liver injury through the inhibition of endoplasmic reticulum stress-mediated ferroptosis by
activating the AMPK/SIRT1 pathway. Ecotoxicol. Environ. Saf. 2023,262, 115331. [CrossRef] [PubMed]
63.
Zhang, M.; Cui, J.; Chen, H.; Wang, Y.; Kuai, X.; Sun, S.; Tang, Q.; Zong, F.; Chen, Q.; Wu, J.; et al. High-Dosage NMN Promotes
Ferroptosis to Suppress Lung Adenocarcinoma Growth through the NAM-Mediated SIRT1-AMPK-ACC Pathway. Cancers 2023,
15, 2427. [CrossRef] [PubMed]
64.
Siblini, Y.; Namour, F.; Oussalah, A.; Gueant, J.L.; Chery, C. Stemness of Normal and Cancer Cells: The Influence of Methionine
Needs and SIRT1/PGC-1alpha/PPAR-alpha Players. Cells 2022,11, 3607. [CrossRef] [PubMed]
65.
Gurd, B.J.; Menezes, E.S.; Arhen, B.B.; Islam, H. Impacts of altered exercise volume, intensity, and duration on the activation of
AMPK and CaMKII and increases in PGC-1alpha mRNA. Semin. Cell Dev. Biol. 2023,143, 17–27. [CrossRef] [PubMed]
Antioxidants 2024,13, 395 24 of 31
66.
Liao, J.; Xie, X.; Wang, N.; Wang, Y.; Zhao, J.; Chen, F.; Qu, F.; Wen, W.; Miao, J.; Cui, H. Formononetin promotes fatty acid
beta-oxidation to treat non-alcoholic steatohepatitis through SIRT1/PGC-1alpha/PPARalpha pathway. Phytomedicine 2024,124,
155285. [CrossRef] [PubMed]
67.
Shi, D.; Hao, Z.; Qi, W.; Jiang, F.; Liu, K.; Shi, X. Aerobic exercise combined with chlorogenic acid exerts neuroprotective effects and
reverses cognitive decline in Alzheimer’s disease model mice (APP/PS1) via the SIRT1/ /PGC-1alpha/PPARgamma signaling
pathway. Front. Aging Neurosci. 2023,15, 1269952. [CrossRef] [PubMed]
68.
Gu, L.; Ding, X.; Wang, Y.; Gu, M.; Zhang, J.; Yan, S.; Li, N.; Song, Z.; Yin, J.; Lu, L.; et al. Spexin alleviates insulin resistance and
inhibits hepatic gluconeogenesis via the FoxO1/PGC-1alpha pathway in high-fat-diet-induced rats and insulin resistant cells. Int.
J. Biol. Sci. 2019,15, 2815–2829. [CrossRef]
69.
Fan, H.; Sun, Y.; Zhang, X.; Xu, Y.; Ming, Y.; Zhang, L.; Zhao, P. Malvidin promotes PGC-1alpha/Nrf2 signaling to attenuate the
inflammatory response and restore mitochondrial activity in septic acute kidney injury. Chem. Biol. Interact. 2024,388, 110850.
[CrossRef]
70.
Zhang, X.S.; Lu, Y.; Li, W.; Tao, T.; Peng, L.; Wang, W.H.; Gao, S.; Liu, C.; Zhuang, Z.; Xia, D.Y.; et al. Astaxanthin ameliorates
oxidative stress and neuronal apoptosis via SIRT1/NRF2/Prx2/ASK1/p38 after traumatic brain injury in mice. Br. J. Pharmacol.
2021,178, 1114–1132. [CrossRef] [PubMed]
71.
Yang, H.; Gu, Z.T.; Li, L.; Maegele, M.; Zhou, B.Y.; Li, F.; Zhao, M.; Zhao, K.S. SIRT1 plays a neuroprotective role in traumatic
brain injury in rats via inhibiting the p38 MAPK pathway. Acta Pharmacol. Sin. 2017,38, 168–181. [CrossRef] [PubMed]
72.
Zhao, Y.; Luo, P.; Guo, Q.; Li, S.; Zhang, L.; Zhao, M.; Xu, H.; Yang, Y.; Poon, W.; Fei, Z. Interactions between SIRT1 and
MAPK/ERK regulate neuronal apoptosis induced by traumatic brain injury
in vitro
and
in vivo
.Exp. Neurol. 2012,237, 489–498.
[CrossRef] [PubMed]
73.
Ahmadi, A.; Ahrari, S.; Salimian, J.; Salehi, Z.; Karimi, M.; Emamvirdizadeh, A.; Jamalkandi, S.A.; Ghanei, M. p38 MAPK
signaling in chronic obstructive pulmonary disease pathogenesis and inhibitor therapeutics. Cell Commun. Signal 2023,21, 314.
[CrossRef] [PubMed]
74.
Xia, M.; Zhang, Y.; Wu, H.; Zhang, Q.; Liu, Q.; Li, G.; Zhao, T.; Liu, X.; Zheng, S.; Qian, Z.; et al. Forsythoside B attenuates
neuro-inflammation and neuronal apoptosis by inhibition of NF-kappaB and p38-MAPK signaling pathways through activating
Nrf2 post spinal cord injury. Int. Immunopharmacol. 2022,111, 109120. [CrossRef] [PubMed]
75.
Wang, J.; Zhang, Y. Neuroprotective effect of berberine agonist against impairment of learning and memory skills in severe
traumatic brain injury via Sirt1/p38 MAPK expression. Mol. Med. Rep. 2018,17, 6881–6886. [CrossRef] [PubMed]
76.
Li, C.; Wu, Z.; Xue, H.; Gao, Q.; Zhang, Y.; Wang, C.; Zhao, P. Ferroptosis contributes to hypoxic-ischemic brain injury in neonatal
rats: Role of the SIRT1/Nrf2/GPx4 signaling pathway. CNS Neurosci. Ther. 2022,28, 2268–2280. [CrossRef] [PubMed]
77.
Chen, T.; Xu, Y.P.; Chen, Y.; Sun, S.; Yan, Z.Z.; Wang, Y.H. Arc regulates brain damage and neuroinflammation via Sirt1 signaling
following subarachnoid hemorrhage. Brain Res. Bull. 2023,203, 110780. [CrossRef] [PubMed]
78.
Ruan, Q.; Wang, C.; Zhang, Y.; Sun, J. Brevilin A attenuates cartilage destruction in osteoarthritis mouse model by inhibiting
inflammation and ferroptosis via SIRT1/Nrf2/GPX4 signaling pathway. Int. Immunopharmacol. 2023,124, 110924. [CrossRef]
[PubMed]
79.
Zhao, L.; Jin, L.; Yang, B. Saikosaponin A alleviates Staphylococcus aureus-induced mastitis in mice by inhibiting ferroptosis via
SIRT1/Nrf2 pathway. J. Cell Mol. Med. 2023,27, 3443–3450. [CrossRef] [PubMed]
80.
Zhao, L.; Jin, L.; Yang, B. Diosmetin alleviates S. aureus-induced mastitis by inhibiting SIRT1/GPX4 mediated ferroptosis. Life Sci.
2023,331, 122060. [CrossRef] [PubMed]
81.
Chen, Z.; Li, J.; Peng, H.; Zhang, M.; Wu, X.; Gui, F.; Li, W.; Ai, F.; Yu, B.; Liu, Y. Meteorin-like/Meteorin-beta protects LPS-induced
acute lung injury by activating SIRT1-P53-SLC7A11 mediated ferroptosis pathway. Mol. Med. 2023,29, 144. [CrossRef] [PubMed]
82.
Yu, M.; Li, H.; Wang, B.; Wu, Z.; Wu, S.; Jiang, G.; Wang, H.; Huang, Y. Baicalein ameliorates polymyxin B-induced acute renal
injury by inhibiting ferroptosis via regulation of SIRT1/p53 acetylation. Chem. Biol. Interact. 2023,382, 110607. [CrossRef]
[PubMed]
83.
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. Mol. Cell Biochem. 2023,478,
609–620. [CrossRef] [PubMed]
84. Engeland, K. Cell cycle regulation: p53-p21-RB signaling. Cell Death Differ. 2022,29, 946–960. [CrossRef] [PubMed]
85.
Rodencal, J.; Kim, N.; He, A.; Li, V.L.; Lange, M.; He, J.; Tarangelo, A.; Schafer, Z.T.; Olzmann, J.A.; Long, J.Z.; et al. Sensitization
of cancer cells to ferroptosis coincident with cell cycle arrest. Cell Chem. Biol. 2024,31, 234–248 e213. [CrossRef]
86.
Yan, C.; Yang, S.; Shao, S.; Zu, R.; Lu, H.; Chen, Y.; Zhou, Y.; Ying, X.; Xiang, S.; Zhang, P.; et al. Exploring the anti-ferroptosis mech-
anism of Kai-Xin-San against Alzheimer’s disease through integrating network pharmacology, bioinformatics, and experimental
validation strategy in vivo and in vitro. J. Ethnopharmacol. 2024,326, 117915. [CrossRef] [PubMed]
87.
Yuan, B.; Zhao, X.D.; Shen, J.D.; Chen, S.J.; Huang, H.Y.; Zhou, X.M.; Han, Y.L.; Zhou, L.J.; Lu, X.J.; Wu, Q. Activation of SIRT1
Alleviates Ferroptosis in the Early Brain Injury after Subarachnoid Hemorrhage. Oxidative Med. Cell. Longev. 2022,2022, 9069825.
[CrossRef] [PubMed]
88.
Zhang, J.; Zhu, Q.; Peng, Z.; Li, X.J.; Ding, P.F.; Gao, S.; Sheng, B.; Liu, Y.; Lu, Y.; Zhuang, Z.; et al. Menaquinone-4 attenuates
ferroptosis by upregulating DHODH through activation of SIRT1 after subarachnoid hemorrhage. Free Radic. Biol. Med. 2024,210,
416–429. [CrossRef] [PubMed]
Antioxidants 2024,13, 395 25 of 31
89.
Peng, X.; Wang, J.; Peng, J.; Jiang, H.; Le, K. Resveratrol Improves Synaptic Plasticity in Hypoxic-Ischemic Brain Injury in Neonatal
Mice via Alleviating SIRT1/NF-kappaB Signaling-Mediated Neuroinflammation. J. Mol. Neurosci. 2022,72, 113–125. [CrossRef]
[PubMed]
90.
Chen, X.; Chen, C.; Fan, S.; Wu, S.; Yang, F.; Fang, Z.; Fu, H.; Li, Y. Omega-3 polyunsaturated fatty acid attenuates the inflammatory
response by modulating microglia polarization through SIRT1-mediated deacetylation of the HMGB1/NF-kappaB pathway
following experimental traumatic brain injury. J. Neuroinflamm. 2018,15, 116. [CrossRef] [PubMed]
91.
Chen, Y.; Peng, F.; Yang, C.; Hou, H.; Xing, Z.; Chen, J.; Liu, L.; Peng, C.; Li, D. SIRT1 activation by 2,3,5,6-tetramethylpyrazine
alleviates neuroinflammation via inhibiting M1 microglia polarization. Front. Immunol. 2023,14, 1206513. [CrossRef] [PubMed]
92.
Zhao, Y.S.; Li, J.Y.; Li, Z.C.; Wang, L.L.; Gan, C.L.; Chen, J.; Jiang, S.Y.; Aschner, M.; Ou, S.Y.; Jiang, Y.M. Sodium Para-aminosalicylic
Acid Inhibits Lead-Induced Neuroinflammation in Brain Cortex of Rats by Modulating SIRT1/HMGB1/NF-kappaB Pathway.
Neurochem. Res. 2023,48, 238–249. [CrossRef] [PubMed]
93.
Zhang, W.; Qian, S.; Tang, B.; Kang, P.; Zhang, H.; Shi, C. Resveratrol inhibits ferroptosis and decelerates heart failure progression
via Sirt1/p53 pathway activation. J. Cell Mol. Med. 2023,27, 3075–3089. [CrossRef] [PubMed]
94.
Zou, P.; Liu, X.; Li, G.; Wang, Y. Resveratrol pretreatment attenuates traumatic brain injury in rats by suppressing NLRP3
inflammasome activation via SIRT1. Mol. Med. Rep. 2018,17, 3212–3217. [CrossRef] [PubMed]
95.
Grabowska, A.D.; Watroba, M.; Witkowska, J.; Mikulska, A.; Sepulveda, N.; Szukiewicz, D. Interplay between Systemic Glycemia
and Neuroprotective Activity of Resveratrol in Modulating Astrocyte SIRT1 Response to Neuroinflammation. Int. J. Mol. Sci.
2023,24, 11640. [CrossRef] [PubMed]
96.
Yan, R.; Lin, B.; Jin, W.; Tang, L.; Hu, S.; Cai, R. NRF2, a Superstar of Ferroptosis. Antioxidants 2023,12, 1739. [CrossRef] [PubMed]
97.
Martin, D.; Rojo, A.I.; Salinas, M.; Diaz, R.; Gallardo, G.; Alam, J.; De Galarreta, C.M.; Cuadrado, A. Regulation of heme
oxygenase-1 expression through the phosphatidylinositol 3-kinase/Akt pathway and the Nrf2 transcription factor in response to
the antioxidant phytochemical carnosol. J. Biol. Chem. 2004,279, 8919–8929. [CrossRef] [PubMed]
98.
Xie, X.; Wang, F.; Ge, W.; Meng, X.; Fan, L.; Zhang, W.; Wang, Z.; Ding, M.; Gu, S.; Xing, X.; et al. Scutellarin attenuates oxidative
stress and neuroinflammation in cerebral ischemia/reperfusion injury through PI3K/Akt-mediated Nrf2 signaling pathways.
Eur. J. Pharmacol. 2023,957, 175979. [CrossRef]
99.
Cheng, Y.; Gao, Y.; Li, J.; Rui, T.; Li, Q.; Chen, H.; Jia, B.; Song, Y.; Gu, Z.; Wang, T.; et al. TrkB agonist N-acetyl serotonin promotes
functional recovery after traumatic brain injury by suppressing ferroptosis via the PI3K/Akt/Nrf2/Ferritin H pathway. Free
Radic. Biol. Med. 2023,194, 184–198. [CrossRef] [PubMed]
100.
Wu, T.; Zhao, F.; Gao, B.; Tan, C.; Yagishita, N.; Nakajima, T.; Wong, P.K.; Chapman, E.; Fang, D.; Zhang, D.D. Hrd1 suppresses
Nrf2-mediated cellular protection during liver cirrhosis. Genes. Dev. 2014,28, 708–722. [CrossRef]
101.
Feng, X.; Ma, W.; Zhu, J.; Jiao, W.; Wang, Y. Dexmedetomidine alleviates early brain injury following traumatic brain injury by
inhibiting autophagy and neuroinflammation through the ROS/Nrf2 signaling pathway. Mol. Med. Rep. 2021,24, 661. [CrossRef]
[PubMed]
102.
Wang, C.; Chen, S.; Guo, H.; Jiang, H.; Liu, H.; Fu, H.; Wang, D. Forsythoside A Mitigates Alzheimer’s-like Pathology by Inhibiting
Ferroptosis-mediated Neuroinflammation via Nrf2/GPX4 Axis Activation. Int. J. Biol. Sci. 2022,18, 2075–2090. [CrossRef]
[PubMed]
103.
Lee, J.; Hyun, D.H. The Interplay between Intracellular Iron Homeostasis and Neuroinflammation in Neurodegenerative Diseases.
Antioxidants 2023,12, 918. [CrossRef] [PubMed]
104.
Cheng, H.; Wang, P.; Wang, N.; Dong, W.; Chen, Z.; Wu, M.; Wang, Z.; Yu, Z.; Guan, D.; Wang, L.; et al. Neuroprotection of NRF2
against Ferroptosis after Traumatic Brain Injury in Mice. Antioxidants 2023,12, 731. [CrossRef] [PubMed]
105.
Pietsch, E.C.; Chan, J.Y.; Torti, F.M.; Torti, S.V. Nrf2 mediates the induction of ferritin H in response to xenobiotics and cancer
chemopreventive dithiolethiones. J. Biol. Chem. 2003,278, 2361–2369. [CrossRef] [PubMed]
106.
Zhang, Y.; Lan, J.; Zhao, D.; Ruan, C.; Zhou, J.; Tan, H.; Bao, Y. Netrin-1 upregulates GPX4 and prevents ferroptosis after traumatic
brain injury via the UNC5B/Nrf2 signaling pathway. CNS Neurosci. Ther. 2023,29, 216–227. [CrossRef] [PubMed]
107.
Ma, X.Z.; Chen, L.L.; Qu, L.; Li, H.; Wang, J.; Song, N.; Xie, J.X. Gut microbiota-induced CXCL1 elevation triggers early
neuroinflammation in the substantia nigra of Parkinsonian mice. Acta Pharmacol. Sin. 2024,45, 52–65. [CrossRef]
108.
Duan, C.; Jiao, D.; Wang, H.; Wu, Q.; Men, W.; Yan, H.; Li, C. Activation of the PPARgamma Prevents Ferroptosis-Induced
Neuronal Loss in Response to Intracerebral Hemorrhage Through Synergistic Actions With the Nrf2. Front. Pharmacol. 2022,13,
869300. [CrossRef] [PubMed]
109.
Lv, J.; Xu, S.; Meng, C.; Wang, Y.; Ji, L.; Li, X.; Wang, X.; Li, Q. Ferroptosis participated in hippocampal neuroinflammation damage
of in offspring rats after maternal sleep deprivation. J. Neuroimmunol. 2023,375, 578021. [CrossRef] [PubMed]
110.
Cui, Y.; Zhang, Z.; Zhou, X.; Zhao, Z.; Zhao, R.; Xu, X.; Kong, X.; Ren, J.; Yao, X.; Wen, Q.; et al. Microglia and macrophage
exhibit attenuated inflammatory response and ferroptosis resistance after RSL3 stimulation via increasing Nrf2 expression. J.
Neuroinflammation 2021,18, 249. [CrossRef] [PubMed]
111.
Gao, Q.; Zhao, Y.; Luo, R.; Su, M.; Zhang, C.; Li, C.; Liu, B.; Zhou, X. Intrathecal umbilical cord mesenchymal stem cells
injection alleviates neuroinflammation and oxidative stress in the cyclophosphamide-induced interstitial cystitis rats through the
Sirt1/Nrf2/HO-1 pathway. Life Sci. 2023,331, 122045. [CrossRef] [PubMed]
Antioxidants 2024,13, 395 26 of 31
112.
Zhao, X.J.; Zhu, H.Y.; Wang, X.L.; Lu, X.W.; Pan, C.L.; Xu, L.; Liu, X.; Xu, N.; Zhang, Z.Y. Oridonin Ameliorates Traumatic
Brain Injury-Induced Neurological Damage by Improving Mitochondrial Function and Antioxidant Capacity and Suppressing
Neuroinflammation through the Nrf2 Pathway. J. Neurotrauma 2022,39, 530–543. [CrossRef] [PubMed]
113.
Wang, L.; Ding, Y.Y.; Wu, Y.Q.; Zhao, C.; Wu, J.; Wang, W.J.; Meng, F.H. Koumine ameliorates neuroinflammation by regulating
microglia polarization via activation of Nrf2/HO-1 pathway. Biomed. Pharmacother. 2023,167, 115608. [CrossRef] [PubMed]
114.
Guo, Y.; Ou, C.; Zhang, N.; Liu, Q.; Xiong, K.; Yu, J.; Cheng, H.; Chen, L.; Ma, M.; Xu, J.; et al. Roflumilast attenuates
neuroinflammation post retinal ischemia/reperfusion injury by regulating microglia phenotype via the Nrf2/STING/NF-kappaB
pathway. Int. Immunopharmacol. 2023,124, 110952. [CrossRef]
115.
Xiaodan, S.; Peiyan, Z.; Hui, L.; Yan, L.; Ying, C. NRF2 participates in the suppressive tumor immune microenvironment of
KRAS/KEAP1 co-mutant non-small cell lung cancer by inhibiting the STING pathway. Genes. Dis. 2023,10, 1727–1730. [CrossRef]
[PubMed]
116.
Yang, T.; Qu, X.; Zhao, J.; Wang, X.; Wang, Q.; Dai, J.; Zhu, C.; Li, J.; Jiang, L. Macrophage PTEN controls STING-induced
inflammation and necroptosis through NICD/NRF2 signaling in APAP-induced liver injury. Cell Commun. Signal. 2023,21, 160.
[CrossRef] [PubMed]
117.
He, L.; Liu, D.; Zhou, W.; Han, Y.; Ju, Y.; Liu, H.; Chen, Y.; Yu, J.; Wang, L.; Wang, J.; et al. The innate immune sensor STING
accelerates neointima formation via NF-kappaB signaling pathway. Int. Immunopharmacol. 2023,121, 110412. [CrossRef]
118.
Bellezza, I.; Tucci, A.; Galli, F.; Grottelli, S.; Mierla, A.L.; Pilolli, F.; Minelli, A. Inhibition of NF-kappaB nuclear translocation via
HO-1 activation underlies alpha-tocopheryl succinate toxicity. J. Nutr. Biochem. 2012,23, 1583–1591. [CrossRef] [PubMed]
119.
Gao, Y.; Zhang, H.; Wang, J.; Li, F.; Li, X.; Li, T.; Wang, C.; Li, L.; Peng, R.; Liu, L.; et al. Annexin A5 ameliorates traumatic brain
injury-induced neuroinflammation and neuronal ferroptosis by modulating the NF-kB/HMGB1 and Nrf2/HO-1 pathways. Int.
Immunopharmacol. 2023,114, 109619. [CrossRef] [PubMed]
120.
Capece, D.; Verzella, D.; Flati, I.; Arboretto, P.; Cornice, J.; Franzoso, G. NF-kappaB: Blending metabolism, immunity, and
inflammation. Trends Immunol. 2022,43, 757–775. [CrossRef] [PubMed]
121.
Singh, S.S.; Rai, S.N.; Birla, H.; Zahra, W.; Rathore, A.S.; Singh, S.P. NF-kappaB-Mediated Neuroinflammation in Parkinson’s
Disease and Potential Therapeutic Effect of Polyphenols. Neurotox. Res. 2020,37, 491–507. [CrossRef] [PubMed]
122.
Ji, H.Z.; Chen, L.; Ren, M.; Li, S.; Liu, T.Y.; Chen, H.J.; Yu, H.H.; Sun, Y. CXCL8 Promotes Endothelial-to-Mesenchymal Transition
of Endothelial Cells and Protects Cells from Erastin-Induced Ferroptosis via CXCR2-Mediated Activation of the NF-kappaB
Signaling Pathway. Pharmaceuticals 2023,16, 1210. [CrossRef] [PubMed]
123.
Tamiya, H.; Urushihara, N.; Shizuma, K.; Ogawa, H.; Nakai, S.; Wakamatsu, T.; Takenaka, S.; Kakunaga, S. SHARPIN Enhances
Ferroptosis in Synovial Sarcoma Cells via NF-kappaB- and PRMT5-Mediated PGC1alpha Reduction. Cancers 2023,15, 3484.
[CrossRef] [PubMed]
124.
Kaur, U.; Banerjee, P.; Bir, A.; Sinha, M.; Biswas, A.; Chakrabarti, S. Reactive oxygen species, redox signaling and neuroinflamma-
tion in Alzheimer’s disease: The NF-kappaB connection. Curr. Top. Med. Chem. 2015,15, 446–457. [CrossRef] [PubMed]
125.
Chen, P.H.; Wu, J.; Ding, C.C.; Lin, C.C.; Pan, S.; Bossa, N.; Xu, Y.; Yang, W.H.; Mathey-Prevot, B.; Chi, J.T. Kinome screen of
ferroptosis reveals a novel role of ATM in regulating iron metabolism. Cell Death Differ. 2020,27, 1008–1022. [CrossRef] [PubMed]
126.
Jia, H.; Liu, X.; Cao, Y.; Niu, H.; Lan, Z.; Li, R.; Li, F.; Sun, D.; Shi, M.; Wa, L.; et al. Deferoxamine ameliorates neurological
dysfunction by inhibiting ferroptosis and neuroinflammation after traumatic brain injury. Brain Res. 2023,1812, 148383. [CrossRef]
127. Hambright, W.S.; Fonseca, R.S.; Chen, L.; Na, R.; Ran, Q. Ablation of ferroptosis regulator glutathione peroxidase 4 in forebrain
neurons promotes cognitive impairment and neurodegeneration. Redox Biol. 2017,12, 8–17. [CrossRef] [PubMed]
128.
Ren, Y.; Zhu, D.; Han, X.; Zhang, Q.; Chen, B.; Zhou, P.; Wei, Z.; Zhang, Z.; Cao, Y.; Zou, H. HMGB1: A double-edged sword and
therapeutic target in the female reproductive system. Front. Immunol. 2023,14, 1238785. [CrossRef] [PubMed]
129.
Chen, Y.; He, W.; Wei, H.; Chang, C.; Yang, L.; Meng, J.; Long, T.; Xu, Q.; Zhang, C. Srs11-92, a ferrostatin-1 analog, improves
oxidative stress and neuroinflammation via Nrf2 signal following cerebral ischemia/reperfusion injury. CNS Neurosci. Ther. 2023,
29, 1667–1677. [CrossRef] [PubMed]
130.
Lee, S.; Ju, I.G.; Choi, Y.; Park, S.; Oh, M.S. Trichosanthis Semen Suppresses Lipopolysaccharide-Induced Neuroinflammation by
Regulating the NF-kappaB Signaling Pathway and HO-1 Expression in Microglia. Toxins 2021,13, 898. [CrossRef] [PubMed]
131.
Zhang, S.S.; Liu, M.; Liu, D.N.; Shang, Y.F.; Wang, Y.H.; Du, G.H. ST2825, a Small Molecule Inhibitor of MyD88, Suppresses
NF-kappaB Activation and the ROS/NLRP3/Cleaved Caspase-1 Signaling Pathway to Attenuate Lipopolysaccharide-Stimulated
Neuroinflammation. Molecules 2022,27, 2990. [CrossRef] [PubMed]
132.
Guan, F.; Zhou, X.; Li, P.; Wang, Y.; Liu, M.; Li, F.; Cui, Y.; Huang, T.; Yao, M.; Zhang, Y.; et al. MG53 attenuates lipopolysaccharide-
induced neurotoxicity and neuroinflammation via inhibiting TLR4/NF-kappaB pathway
in vitro
and
in vivo
.Prog. Neuropsy-
chopharmacol. Biol. Psychiatry 2019,95, 109684. [CrossRef] [PubMed]
133.
An, Q.; Xia, J.; Pu, F.; Shi, S. MCPIP1 alleviates depressive-like behaviors in mice by inhibiting the TLR4/TRAF6/NF-kappaB
pathway to suppress neuroinflammation. Mol. Med. Rep. 2024,29, 6. [CrossRef] [PubMed]
134.
Xu, Y.; Zhang, J.; Gao, F.; Cheng, W.; Zhang, Y.; Wei, C.; Zhang, S.; Gao, X. Engeletin alleviates cerebral ischemia reperfusion-
induced neuroinflammation via the HMGB1/TLR4/NF-kappaB network. J. Cell Mol. Med. 2023,27, 1653–1663. [CrossRef]
[PubMed]
Antioxidants 2024,13, 395 27 of 31
135.
Zhang, Y.; Ye, P.; Zhu, H.; Gu, L.; Li, Y.; Feng, S.; Zeng, Z.; Chen, Q.; Zhou, B.; Xiong, X. Neutral polysaccharide from Gastrodia
elata alleviates cerebral ischemia-reperfusion injury by inhibiting ferroptosis-mediated neuroinflammation via the NRF2/HO-1
signaling pathway. CNS Neurosci. Ther. 2023,30, e14456. [CrossRef] [PubMed]
136.
Wang, M.; Tang, G.; Zhou, C.; Guo, H.; Hu, Z.; Hu, Q.; Li, G. Revisiting the intersection of microglial activation and neuroin-
flammation in Alzheimer’s disease from the perspective of ferroptosis. Chem. Biol. Interact. 2023,375, 110387. [CrossRef]
[PubMed]
137.
Pozniak, P.D.; White, M.K.; Khalili, K. TNF-alpha/NF-kappaB signaling in the CNS: Possible connection to EPHB2. J. Neuroimmune
Pharmacol. 2014,9, 133–141. [CrossRef] [PubMed]
138.
Wang, X.; Li, S.; Yu, J.; Wang, W.; Du, Z.; Gao, S.; Ma, Y.; Tang, R.; Liu, T.; Ma, S.; et al. Saikosaponin B2 ameliorates depression-
induced microglia activation by inhibiting ferroptosis-mediated neuroinflammation and ER stress. J. Ethnopharmacol. 2023,316,
116729. [CrossRef] [PubMed]
139.
Abdulaal, W.H.; Omar, U.M.; Zeyadi, M.; El-Agamy, D.S.; Alhakamy, N.A.; Ibrahim, S.R.M.; Almalki, N.A.R.; Asfour, H.Z.; Al-
Rabia, M.W.; Mohamed, G.A.; et al. Modulation of the crosstalk between Keap1/Nrf2/HO-1 and NF-kappaB signaling pathways
by Tomatidine protects against inflammation/oxidative stress-driven fulminant hepatic failure in mice. Int. Immunopharmacol.
2024,130, 111732. [CrossRef] [PubMed]
140.
Xiao, X.; Jiang, Y.; Liang, W.; Wang, Y.; Cao, S.; Yan, H.; Gao, L.; Zhang, L. miR-212-5p attenuates ferroptotic neuronal death after
traumatic brain injury by targeting Ptgs2. Mol. Brain 2019,12, 78. [CrossRef] [PubMed]
141.
Zhang, X.H.; Cui, H.; Zheng, S.M.; Lu, Y.; Yuan, H.W.; Zhang, L.; Wang, H.H.; Du, R.S. Electroacupuncture regulates microglial
polarization via inhibiting NF-kappaB/COX2 pathway following traumatic brain injury. Brain Res. 2023,1818, 148516. [CrossRef]
[PubMed]
142.
Prabhakaran, J.; Molotkov, A.; Mintz, A.; Mann, J.J. Progress in PET Imaging of Neuroinflammation Targeting COX-2 Enzyme.
Molecules 2021,26, 3208. [CrossRef] [PubMed]
143.
Hakan, T.; Toklu, H.Z.; Biber, N.; Ozevren, H.; Solakoglu, S.; Demirturk, P.; Aker, F.V. Effect of COX-2 inhibitor meloxicam against
traumatic brain injury-induced biochemical, histopathological changes and blood-brain barrier permeability. Neurol. Res. 2010,
32, 629–635. [CrossRef] [PubMed]
144.
Liu, J.; Zhou, Y.; Xie, C.; Li, C.; Ma, L.; Zhang, Y. Anti-Ferroptotic Effects of bone Marrow Mesenchymal Stem Cell-Derived
Extracellular Vesicles Loaded with Ferrostatin-1 in Cerebral ischemia-reperfusion Injury Associate with the GPX4/COX-2 Axis.
Neurochem. Res. 2023,48, 502–518. [CrossRef]
145.
Moussa, N.; Dayoub, N. Exploring the role of COX-2 in Alzheimer’s disease: Potential therapeutic implications of COX-2
inhibitors. Saudi Pharm. J. 2023,31, 101729. [CrossRef] [PubMed]
146.
Niranjan, R.; Mishra, K.P.; Thakur, A.K. Inhibition of Cyclooxygenase-2 (COX-2) Initiates Autophagy and Potentiates MPTP-
Induced Autophagic Cell Death of Human Neuroblastoma Cells, SH-SY5Y: An Inside in the Pathology of Parkinson’s Disease.
Mol. Neurobiol. 2018,55, 8038–8050. [CrossRef] [PubMed]
147.
Ayala, A.; Munoz, M.F.; Arguelles, S. Lipid peroxidation: Production, metabolism, and signaling mechanisms of malondialdehyde
and 4-hydroxy-2-nonenal. Oxidative Med. Cell. Longev. 2014,2014, 360438. [CrossRef] [PubMed]
148.
Song, H.; Park, J.; Bui, P.T.C.; Choi, K.; Gye, M.C.; Hong, Y.C.; Kim, J.H.; Lee, Y.J. Bisphenol A induces COX-2 through the
mitogen-activated protein kinase pathway and is associated with levels of inflammation-related markers in elderly populations.
Environ. Res. 2017,158, 490–498. [CrossRef]
149.
Cherukuri, D.P.; Goulet, A.C.; Inoue, H.; Nelson, M.A. Selenomethionine regulates cyclooxygenase-2 (COX-2) expression through
nuclear factor-kappa B (NF-kappaB) in colon cancer cells. Cancer Biol. Ther. 2005,4, 175–180. [CrossRef]
150.
Chen, Y.F.; Wang, Y.W.; Huang, W.S.; Lee, M.M.; Wood, W.G.; Leung, Y.M.; Tsai, H.Y. Trans-Cinnamaldehyde, An Essential
Oil in Cinnamon Powder, Ameliorates Cerebral Ischemia-Induced Brain Injury via Inhibition of Neuroinflammation Through
Attenuation of iNOS, COX-2 Expression and NFkappa-B Signaling Pathway. Neuromol. Med. 2016,18, 322–333. [CrossRef]
151.
Yan, J.J.; Du, G.H.; Qin, X.M.; Gao, L. Baicalein attenuates the neuroinflammation in LPS-activated BV-2 microglial cells through
suppression of pro-inflammatory cytokines, COX2/NF-kappaB expressions and regulation of metabolic abnormality. Int.
Immunopharmacol. 2020,79, 106092. [CrossRef]
152.
Haile, M.; Boutajangout, A.; Chung, K.; Chan, J.; Stolper, T.; Vincent, N.; Batchan, M.; D’Urso, J.; Lin, Y.; Kline, R.; et al. The Cox-2
Inhibitor Meloxicam Ameliorates Neuroinflammation and Depressive Behavior in Adult Mice after Splenectomy. J. Neurophysiol.
Neurol. Disord. 2016,3, 101.
153.
Zhou, Y.; Zhou, H.; Hua, L.; Hou, C.; Jia, Q.; Chen, J.; Zhang, S.; Wang, Y.; He, S.; Jia, E. Verification of ferroptosis and pyroptosis
and identification of PTGS2 as the hub gene in human coronary artery atherosclerosis. Free Radic. Biol. Med. 2021,171, 55–68.
[CrossRef] [PubMed]
154.
Cao, Y.; Li, Y.; He, C.; Yan, F.; Li, J.R.; Xu, H.Z.; Zhuang, J.F.; Zhou, H.; Peng, Y.C.; Fu, X.J.; et al. Selective Ferroptosis Inhibitor
Liproxstatin-1 Attenuates Neurological Deficits and Neuroinflammation After Subarachnoid Hemorrhage. Neurosci. Bull. 2021,
37, 535–549. [CrossRef] [PubMed]
155.
Zou, C.; Xu, F.; Shen, J.; Xu, S. Identification of a Ferroptosis-Related Prognostic Gene PTGS2 Based on Risk Modeling and
Immune Microenvironment of Early-Stage Cervical Cancer. J. Oncol. 2022,2022, 3997562. [CrossRef] [PubMed]
156.
Zhang, T.; Yang, M.; Ma, C.; Wei, X.; Zhang, Z. BACH1 encourages ferroptosis by activating KDM4C-mediated COX2 demethyla-
tion after cerebral ischemia-reperfusion injury. Eur. J. Neurosci. 2023,58, 2194–2214. [CrossRef] [PubMed]
Antioxidants 2024,13, 395 28 of 31
157.
Sun, Y.; Chen, P.; Zhai, B.; Zhang, M.; Xiang, Y.; Fang, J.; Xu, S.; Gao, Y.; Chen, X.; Sui, X.; et al. The emerging role of ferroptosis in
inflammation. Biomed. Pharmacother. 2020,127, 110108. [CrossRef] [PubMed]
158.
Liu, T.; Shu, J.; Liu, Y.; Xie, J.; Li, T.; Li, H.; Li, L. Atorvastatin attenuatesferroptosis-dependent myocardial injury and inflammation
following coronary microembolization via the Hif1a/Ptgs2 pathway. Front. Pharmacol. 2022,13, 1057583. [CrossRef]
159.
Gao, F.; Zhao, Y.; Zhang, B.; Xiao, C.; Sun, Z.; Gao, Y.; Dou, X. Suppression of lncRNA Gm47283 attenuates myocardial infarction
via miR-706/ Ptgs2/ferroptosis axis. Bioengineered 2022,13, 10786–10802. [CrossRef] [PubMed]
160.
Liu, C.; Lu, J.; Yuan, T.; Xie, L.; Zhang, L. EPC-exosomal miR-26a-5p improves airway remodeling in COPD by inhibiting
ferroptosis of bronchial epithelial cells via PTGS2/PGE2 signaling pathway. Sci. Rep. 2023,13, 6126. [CrossRef]
161.
Xue, Z.; Zhang, Z.; Liu, H.; Li, W.; Guo, X.; Zhang, Z.; Liu, Y.; Jia, L.; Li, Y.; Ren, Y.; et al. lincRNA-Cox2 regulates NLRP3
inflammasome and autophagy mediated neuroinflammation. Cell Death Differ. 2019,26, 130–145. [CrossRef] [PubMed]
162.
Jiang, H.; Sun, Z.; Zhu, X.; Li, F.; Chen, Q. Essential genes Ptgs2, Tlr4, and Ccr2 regulate neuro-inflammation during the acute
phase of cerebral ischemic in mice. Sci. Rep. 2023,13, 13021. [CrossRef] [PubMed]
163.
Ma, P.; Li, Y.; Zhang, W.; Fang, F.; Sun, J.; Liu, M.; Li, K.; Dong, L. Long Non-coding RNA MALAT1 Inhibits Neuron Apoptosis
and Neuroinflammation While Stimulates Neurite Outgrowth and Its Correlation With MiR-125b Mediates PTGS2, CDK5 and
FOXQ1 in Alzheimer’s Disease. Curr. Alzheimer Res. 2019,16, 596–612. [CrossRef] [PubMed]
164.
Yang, H.; Wang, H.; Shu, Y.; Li, X. miR-103 Promotes Neurite Outgrowth and Suppresses Cells Apoptosis by Targeting
Prostaglandin-Endoperoxide Synthase 2 in Cellular Models of Alzheimer’s Disease. Front. Cell Neurosci. 2018,12, 91. [CrossRef]
[PubMed]
165.
Li, Y.; Wang, J.; Chen, S.; Wu, P.; Xu, S.; Wang, C.; Shi, H.; Bihl, J. miR-137 boosts the neuroprotective effect of endothelial
progenitor cell-derived exosomes in oxyhemoglobin-treated SH-SY5Y cells partially via COX2/PGE2 pathway. Stem Cell Res.
Ther. 2020,11, 330. [CrossRef] [PubMed]
166.
Yi, T.T.; Zhang, L.M.; Huang, X.N. Glycyrrhizic acid protects against temporal lobe epilepsy in young rats by regulating neuronal
ferroptosis through the miR-194-5p/PTGS2 axis. Kaohsiung J. Med. Sci. 2023,39, 154–165. [CrossRef] [PubMed]
167.
Kim, M.E.; Jung, I.; Na, J.Y.; Lee, Y.; Lee, J.; Lee, J.S.; Lee, J.S. Pseudane-VII Regulates LPS-Induced Neuroinflammation in Brain
Microglia Cells through the Inhibition of iNOS Expression. Molecules 2018,23, 3196. [CrossRef] [PubMed]
168.
Hu, C.; He, M.; Chen, M.; Xu, Q.; Li, S.; Cui, Y.; Qiu, X.; Tian, W. Amelioration of Neuropathic Pain and Attenuation of
Neuroinflammation Responses by Tetrahydropalmatine Through the p38MAPK/NF-kappaB/iNOS Signaling Pathways in
Animal and Cellular Models. Inflammation 2022,45, 891–903. [CrossRef]
169.
Wang, S.S.; Jia, J.; Wang, Z. Mesenchymal Stem Cell-Derived Extracellular Vesicles Suppresses iNOS Expression and Ameliorates
Neural Impairment in Alzheimer’s Disease Mice. J. Alzheimers Dis. 2018,61, 1005–1013. [CrossRef] [PubMed]
170.
Dong, Y.; Han, L.L.; Xu, Z.X. Suppressed microRNA-96 inhibits iNOS expression and dopaminergic neuron apoptosis through
inactivating the MAPK signaling pathway by targeting CACNG5 in mice with Parkinson’s disease. Mol. Med. 2018,24, 61.
[CrossRef] [PubMed]
171.
Zhou, M.; Wang, C.M.; Yang, W.L.; Wang, P. Microglial CD14 activated by iNOS contributes to neuroinflammation in cerebral
ischemia. Brain Res. 2013,1506, 105–114. [CrossRef] [PubMed]
172.
Zheng, S.; Wang, C.; Lin, L.; Mu, S.; Liu, H.; Hu, X.; Chen, X.; Wang, S. TNF-alpha Impairs Pericyte-Mediated Cerebral
Microcirculation via the NF-kappaB/iNOS Axis after Experimental Traumatic Brain Injury. J. Neurotrauma 2023,40, 349–364.
[CrossRef] [PubMed]
173.
Adeoluwa, O.A.; Olayinka, J.N.; Adeoluwa, G.O.; Akinluyi, E.T.; Adeniyi, F.R.; Fafure, A.; Nebo, K.; Edem, E.E.; Eduviere, A.T.;
Abubakar, B. Quercetin abrogates lipopolysaccharide-induced depressive-like symptoms by inhibiting neuroinflammation via
microglial NLRP3/NFkappaB/iNOS signaling pathway. Behav. Brain Res. 2023,450, 114503. [CrossRef] [PubMed]
174.
Gong, L.; Zhu, T.; Chen, C.; Xia, N.; Yao, Y.; Ding, J.; Xu, P.; Li, S.; Sun, Z.; Dong, X.; et al. Miconazole exerts disease-modifying
effects during epilepsy by suppressing neuroinflammation via NF-kappaB pathway and iNOS production. Neurobiol. Dis. 2022,
172, 105823. [CrossRef] [PubMed]
175. Urrutia, P.J.; Borquez, D.A.; Nunez, M.T. Inflaming the Brain with Iron. Antioxidants 2021,10, 61. [CrossRef] [PubMed]
176.
Ko, C.J.; Gao, S.L.; Lin, T.K.; Chu, P.Y.; Lin, H.Y. Ferroptosis as a Major Factor and Therapeutic Target for Neuroinflammation in
Parkinson’s Disease. Biomedicines 2021,9, 1679. [CrossRef] [PubMed]
177.
Qu, W.; Cheng, Y.; Peng, W.; Wu, Y.; Rui, T.; Luo, C.; Zhang, J. Targeting iNOS Alleviates Early Brain Injury After Experimental
Subarachnoid Hemorrhage via Promoting Ferroptosis of M1 Microglia and Reducing Neuroinflammation. Mol. Neurobiol. 2022,
59, 3124–3139. [CrossRef]
178.
Dar, H.H.; Anthonymuthu, T.S.; Ponomareva, L.A.; Souryavong, A.B.; Shurin, G.V.; Kapralov, A.O.; Tyurin, V.A.; Lee, J.S.;
Mallampalli, R.K.; Wenzel, S.E.; et al. A new thiol-independent mechanism of epithelial host defense against Pseudomonas
aeruginosa: iNOS/NO(*) sabotage of theft-ferroptosis. Redox Biol. 2021,45, 102045. [CrossRef]
179.
Mikulska-Ruminska, K.; Anthonymuthu, T.S.; Levkina, A.; Shrivastava, I.H.; Kapralov, A.A.; Bayir, H.; Kagan, V.E.; Bahar, I.
NO(
) Represses the Oxygenation of Arachidonoyl PE by 15LOX/PEBP1: Mechanism and Role in Ferroptosis. Int. J. Mol. Sci.
2021,22, 5253. [CrossRef] [PubMed]
180.
Zi, Y.; Wang, X.; Zi, Y.; Yu, H.; Lan, Y.; Fan, Y.; Ren, C.; Liao, K.; Chen, H. Cigarette smoke induces the ROS accumulation and
iNOS activation through deactivation of Nrf-2/SIRT3 axis to mediate the human bronchial epithelium ferroptosis. Free Radic. Biol.
Med. 2023,200, 73–86. [CrossRef]
Antioxidants 2024,13, 395 29 of 31
181.
Wen, Y.; Chen, H.; Zhang, L.; Wu, M.; Zhang, F.; Yang, D.; Shen, J.; Chen, J. Glycyrrhetinic acid induces oxidative/nitrative stress
and drives ferroptosis through activating NADPH oxidases and iNOS, and depriving glutathione in triple-negative breast cancer
cells. Free Radic. Biol. Med. 2021,173, 41–51. [CrossRef]
182.
Kapralov, A.A.; Yang, Q.; Dar, H.H.; Tyurina, Y.Y.; Anthonymuthu, T.S.; Kim, R.; St Croix, C.M.; Mikulska-Ruminska, K.; Liu, B.;
Shrivastava, I.H.; et al. Redox lipid reprogramming commands susceptibility of macrophages and microglia to ferroptotic death.
Nat. Chem. Biol. 2020,16, 278–290. [CrossRef]
183.
Bayir, H.; Kagan, V.E.; Borisenko, G.G.; Tyurina, Y.Y.; Janesko, K.L.; Vagni, V.A.; Billiar, T.R.; Williams, D.L.; Kochanek, P.M.
Enhanced oxidative stress in iNOS-deficient mice after traumatic brain injury: Support for a neuroprotective role of iNOS. J. Cereb.
Blood Flow. Metab. 2005,25, 673–684. [CrossRef]
184.
Chen, X.; Pang, X.; Yeo, A.J.; Xie, S.; Xiang, M.; Shi, B.; Yu, G.; Li, C. The Molecular Mechanisms of Ferroptosis and Its Role in
Blood-Brain Barrier Dysfunction. Front. Cell Neurosci. 2022,16, 889765. [CrossRef]
185.
Tian, D.; Li, W.; Heffron, C.L.; Wang, B.; Mahsoub, H.M.; Sooryanarain, H.; Hassebroek, A.M.; Clark-Deener, S.; LeRoith, T.;
Meng, X.J. Hepatitis E virus infects brain microvascular endothelial cells, crosses the blood-brain barrier, and invades the central
nervous system. Pro. Natl. Acad. Sci. USA 2022,119, e2201862119. [CrossRef]
186.
Huang, Y.-T.; Hsu, Y.-T.; Wu, P.-Y.; Yeh, Y.-M.; Lin, P.-C.; Hsu, K.-F.; Shen, M.-R. Tight junction protein cingulin variant is associated
with cancer susceptibility by overexpressed IQGAP1 and Rac1-dependent epithelial-mesenchymal transition. J. Exp. Clin. Cancer
Res. 2024,43, 65. [CrossRef]
187.
Wang, P.; Jin, L.; Zhang, M.; Wu, Y.; Duan, Z.; Guo, Y.; Wang, C.; Guo, Y.; Chen, W.; Liao, Z.; et al. Blood–brain barrier injury and
neuroinflammation induced by SARS-CoV-2 in a lung–brain microphysiological system. Nat. Biomed. Eng. 2023. [CrossRef]
188.
Li, W.; Tiedt, S.; Lawrence, J.H.; Harrington, M.E.; Musiek, E.S.; Lo, E.H. Circadian Biology and the Neurovascular Unit. Circ. Res.
2024,134, 748–769. [CrossRef]
189.
Zhao, Y.; Liu, Y.; Xu, Y.; Li, K.; Zhou, L.; Qiao, H.; Xu, Q.; Zhao, J. The Role of Ferroptosis in Blood-Brain Barrier Injury. Cell Mol.
Neurobiol. 2023,43, 223–236. [CrossRef]
190.
Wang, P.; Ren, Q.; Shi, M.; Liu, Y.; Bai, H.; Chang, Y.Z. Overexpression of Mitochondrial Ferritin Enhances Blood-Brain Barrier
Integrity Following Ischemic Stroke in Mice by Maintaining Iron Homeostasis in Endothelial Cells. Antioxidants 2022,11, 1257.
[CrossRef]
191.
You, L.; Yu, P.P.; Dong, T.; Guo, W.; Chang, S.; Zheng, B.; Ci, Y.; Wang, F.; Yu, P.; Gao, G.; et al. Astrocyte-derived hepcidin controls
iron traffic at the blood-brain-barrier via regulating ferroportin 1 of microvascular endothelial cells. Cell Death Dis. 2022,13, 667.
[CrossRef]
192.
Rand, D.; Ravid, O.; Atrakchi, D.; Israelov, H.; Bresler, Y.; Shemesh, C.; Omesi, L.; Liraz-Zaltsman, S.; Gosselet, F.; Maskrey,
T.S.; et al. Endothelial Iron Homeostasis Regulates Blood-Brain Barrier Integrity via the HIF2alpha-Ve-Cadherin Pathway.
Pharmaceutics 2021,13, 311. [CrossRef]
193.
Li, J.; Li, M.; Ge, Y.; Chen, J.; Ma, J.; Wang, C.; Sun, M.; Wang, L.; Yao, S.; Yao, C. beta-amyloid protein induces mitophagy-
dependent ferroptosis through the CD36/PINK/PARKIN pathway leading to blood-brain barrier destruction in Alzheimer’s
disease. Cell Biosci. 2022,12, 69. [CrossRef]
194.
Liu, Q.; Song, T.; Chen, B.; Zhang, J.; Li, W. Ferroptosis of brain microvascular endothelial cells contributes to hypoxia-induced
blood-brain barrier injury. FASEB J. 2023,37, e22874. [CrossRef]
195.
Xu, S.; Li, X.; Li, Y.; Li, X.; Lv, E.; Zhang, X.; Shi, Y.; Wang, Y. Neuroprotective effect of Dl-3-n-butylphthalide against ischemia-
reperfusion injury is mediated by ferroptosis regulation via the SLC7A11/GSH/GPX4 pathway and the attenuation of blood-brain
barrier disruption. Front. Aging Neurosci. 2023,15, 1028178. [CrossRef]
196.
Fang, J.; Yuan, Q.; Du, Z.; Fei, M.; Zhang, Q.; Yang, L.; Wang, M.; Yang, W.; Yu, J.; Wu, G.; et al. Ferroptosis in brain microvascular
endothelial cells mediates blood-brain barrier disruption after traumatic brain injury. Biochem. Biophys. Res. Commun. 2022,619,
34–41. [CrossRef]
197.
Meng, S.; Cao, H.; Huang, Y.; Shi, Z.; Li, J.; Wang, Y.; Zhang, Y.; Chen, S.; Shi, H.; Gao, Y. ASK1-K716R reduces neuroinflammation
and white matter injury via preserving blood-brain barrier integrity after traumatic brain injury. J. Neuroinflamm. 2023,20, 244.
[CrossRef]
198.
Villalba, N.; Ma, Y.; Gahan, S.A.; Joly-Amado, A.; Spence, S.; Yang, X.; Nash, K.R.; Yuan, S.Y. Lung infection by Pseudomonas
aeruginosa induces neuroinflammation and blood-brain barrier dysfunction in mice. J. Neuroinflamm. 2023,20, 127. [CrossRef]
199.
Liao, Y.C.; Wang, J.W.; Guo, C.; Bai, M.; Ran, Z.; Wen, L.M.; Ju, B.W.; Ding, Y.; Hu, J.P.; Yang, J.H. Cistanche tubulosa alleviates
ischemic stroke-induced blood-brain barrier damage by modulating microglia-mediated neuroinflammation. J. Ethnopharmacol.
2023,309, 116269. [CrossRef]
200.
Erickson, M.A.; Shulyatnikova, T.; Banks, W.A.; Hayden, M.R. Ultrastructural Remodeling of the Blood-Brain Barrier and
Neurovascular Unit by Lipopolysaccharide-Induced Neuroinflammation. Int. J. Mol. Sci. 2023,24, 1640. [CrossRef]
201.
Larochelle, J.; Yang, C.; Liu, L.; Candelario-Jalil, E. An Unexplored Role for MMP-7 (Matrix Metalloproteinase-7) in Promoting
Gut Permeability After Ischemic Stroke. Stroke 2022,53, 3238–3242. [CrossRef]
202.
Liu, Y.; Mu, Y.; Li, Z.; Yong, V.W.; Xue, M. Extracellular matrix metalloproteinase inducer in brain ischemia and intracerebral
hemorrhage. Front. Immunol. 2022,13, 986469. [CrossRef]
203.
Isik, S.; Yeman Kiyak, B.; Akbayir, R.; Seyhali, R.; Arpaci, T. Microglia Mediated Neuroinflammation in Parkinson’s Disease. Cells
2023,12, 1012. [CrossRef]
Antioxidants 2024,13, 395 30 of 31
204.
Yu, H.; Chang, Q.; Sun, T.; He, X.; Wen, L.; An, J.; Feng, J.; Zhao, Y. Metabolic reprogramming and polarization of microglia in
Parkinson’s disease: Role of inflammasome and iron. Ageing Res. Rev. 2023,90, 102032. [CrossRef]
205.
Liu, S.; Gao, X.; Zhou, S. New Target for Prevention and Treatment of Neuroinflammation: Microglia Iron Accumulation and
Ferroptosis. ASN Neuro 2022,14, 17590914221133236. [CrossRef]
206.
Zhou, X.; Zhao, R.; Lv, M.; Xu, X.; Liu, W.; Li, X.; Gao, Y.; Zhao, Z.; Zhang, Z.; Li, Y.; et al. ACSL4 promotes microglia-mediated
neuroinflammation by regulating lipid metabolism and VGLL4 expression. Brain Behav. Immun. 2023,109, 331–343. [CrossRef]
207.
Wu, Y.; Eisel, U.L.M. Microglia-Astrocyte Communication in Alzheimer’s Disease. J. Alzheimers Dis. 2023,95, 785–803. [CrossRef]
208.
Manu, D.R.; Slevin, M.; Barcutean, L.; Forro, T.; Boghitoiu, T.; Balasa, R. Astrocyte Involvement in Blood-Brain Barrier Function:
A Critical Update Highlighting Novel, Complex, Neurovascular Interactions. Int. J. Mol. Sci. 2023,24, 17146. [CrossRef]
209. Juric, D.M.; Krzan, M.; Lipnik-Stangelj, M. Histamine and astrocyte function. Pharmacol. Res. 2016,111, 774–783. [CrossRef]
210.
Blackburn, D.; Sargsyan, S.; Monk, P.N.; Shaw, P.J. Astrocyte function and role in motor neuron disease: A future therapeutic
target? Glia 2009,57, 1251–1264. [CrossRef]
211.
Murphy-Royal, C.; Ching, S.; Papouin, T. A conceptual framework for astrocyte function. Nat. Neurosci. 2023,26, 1848–1856.
[CrossRef]
212. Verkhratsky, A.; Butt, A.; Li, B.; Illes, P.; Zorec, R.; Semyanov, A.; Tang, Y.; Sofroniew, M.V. Astrocytes in human central nervous
system diseases: A frontier for new therapies. Signal Transduct. Target. Ther. 2023,8, 396. [CrossRef]
213.
Wang, N.; Zhao, Y.; Wu, M.; Li, N.; Yan, C.; Guo, H.; Li, Q.; Li, Q.; Wang, Q. Gemfibrozil Alleviates Cognitive Impairment by
Inhibiting Ferroptosis of Astrocytes via Restoring the Iron Metabolism and Promoting Antioxidant Capacity in Type 2 Diabetes.
Mol. Neurobiol. 2024,61, 1187–1201. [CrossRef]
214.
Xu, S.X.; Xie, X.H.; Yao, L.; Wang, W.; Zhang, H.; Chen, M.M.; Sun, S.; Nie, Z.W.; Nagy, C.; Liu, Z. Human
in vivo
evidence of
reduced astrocyte activation and neuroinflammation in patients with treatment-resistant depression following electroconvulsive
therapy. Psychiatry Clin. Neurosci. 2023,77, 653–664. [CrossRef]
215.
Park, M.W.; Cha, H.W.; Kim, J.; Kim, J.H.; Yang, H.; Yoon, S.; Boonpraman, N.; Yi, S.S.; Yoo, I.D.; Moon, J.S. NOX4 promotes
ferroptosis of astrocytes by oxidative stress-induced lipid peroxidation via the impairment of mitochondrial metabolism in
Alzheimer’s diseases. Redox Biol. 2021,41, 101947. [CrossRef]
216.
Mekhaeil, M.; Conroy, M.J.; Dev, K.K. Elucidating the Therapeutic Utility of Olaparib in Sulfatide-Induced Human Astrocyte
Toxicity and Neuroinflammation. J. Neuroimmune Pharmacol. 2023,18, 592–609. [CrossRef]
217.
Nakano-Kobayashi, A.; Canela, A.; Yoshihara, T.; Hagiwara, M. Astrocyte-targeting therapy rescues cognitive impairment caused
by neuroinflammation via the Nrf2 pathway. Proc. Natl. Acad. Sci. USA 2023,120, e2303809120. [CrossRef]
218.
Liu, C.; Zhao, X.M.; Wang, Q.; Du, T.T.; Zhang, M.X.; Wang, H.Z.; Li, R.P.; Liang, K.; Gao, Y.; Zhou, S.Y.; et al. Astrocyte-derived
SerpinA3N promotes neuroinflammation and epileptic seizures by activating the NF-kappaB signaling pathway in mice with
temporal lobe epilepsy. J. Neuroinflamm. 2023,20, 161. [CrossRef]
219.
Yin, S.; Ma, X.Y.; Sun, Y.F.; Yin, Y.Q.; Long, Y.; Zhao, C.L.; Ma, J.W.; Li, S.; Hu, Y.; Li, M.T.; et al. RGS5 augments astrocyte activation
and facilitates neuroinflammation via TNF signaling. J. Neuroinflamm. 2023,20, 203. [CrossRef]
220.
Davaanyam, D.; Lee, H.; Seol, S.I.; Oh, S.A.; Kim, S.W.; Lee, J.K. HMGB1 induces hepcidin upregulation in astrocytes and causes
an acute iron surge and subsequent ferroptosis in the postischemic brain. Exp. Mol. Med. 2023,55, 2402–2416. [CrossRef]
221.
Ishii, T.; Warabi, E.; Mann, G.E. Circadian control of BDNF-mediated Nrf2 activation in astrocytes protects dopaminergic neurons
from ferroptosis. Free Radic. Biol. Med. 2019,133, 169–178. [CrossRef]
222.
Chen, J.; Qu, X.; Li, Z.; Zhang, D.; Hou, L. Peak Neutrophil-to-Lymphocyte Ratio Correlates with Clinical Outcomes in Patients
with Severe Traumatic Brain Injury. Neurocrit. Care 2019,30, 334–339. [CrossRef]
223.
Li, F.; Weng, G.; Zhou, H.; Zhang, W.; Deng, B.; Luo, Y.; Tao, X.; Deng, M.; Guo, H.; Zhu, S.; et al. The neutrophil-to-
lymphocyte ratio, lymphocyte-to-monocyte ratio, and neutrophil-to-high-density-lipoprotein ratio are correlated with the severity
of Parkinson’s disease. Front. Neurol. 2024,15, 1322228. [CrossRef]
224.
Xu, C.; Cai, L.; Yi, T.; Yi, X.; Hu, Y. Neutrophil-to-lymphocyte ratio is associated with stroke progression and functional outcome
in patients with ischemic stroke. Brain Behav. 2023,13, e3261. [CrossRef]
225.
Nikaido, Y.; Midorikawa, Y.; Furukawa, T.; Shimoyama, S.; Takekawa, D.; Kitayama, M.; Ueno, S.; Kushikata, T.; Hirota, K. The
role of neutrophil gelatinase-associated lipocalin and iron homeostasis in object recognition impairment in aged sepsis-survivor
rats. Sci. Rep. 2022,12, 249. [CrossRef]
226. Liu, Y.W.; Li, S.; Dai, S.S. Neutrophils in traumatic brain injury (TBI): Friend or foe? J. Neuroinflamm. 2018,15, 146. [CrossRef]
227.
Wang, K.; Wang, M.; Liao, X.; Gao, S.; Hua, J.; Wu, X.; Guo, Q.; Xu, W.; Sun, J.; He, Y.; et al. Locally organised and activated
Fth1(hi) neutrophils aggravate inflammation of acute lung injury in an IL-10-dependent manner. Nat. Commun. 2022,13, 7703.
[CrossRef]
228.
Semple, B.D.; Trivedi, A.; Gimlin, K.; Noble-Haeusslein, L.J. Neutrophil elastase mediates acute pathogenesis and is a determinant
of long-term behavioral recovery after traumatic injury to the immature brain. Neurobiol. Dis. 2015,74, 263–280. [CrossRef]
229.
Garcia-Culebras, A.; Duran-Laforet, V.; Pena-Martinez, C.; Ballesteros, I.; Pradillo, J.M.; Diaz-Guzman, J.; Lizasoain, I.; Moro,
M.A. Myeloid cells as therapeutic targets in neuroinflammation after stroke: Specific roles of neutrophils and neutrophil-platelet
interactions. J. Cereb. Blood Flow. Metab. 2018,38, 2150–2164. [CrossRef]
230.
Zhou, J.; Guo, P.; Hao, X.; Sun, X.; Feng, H.; Chen, Z. Neutrophil Extracellular Traps (NETs): A New Therapeutic Target for
Neuroinflammation and Microthrombosis After Subarachnoid Hemorrhage? Transl. Stroke Res. 2023,14, 443–445. [CrossRef]
Antioxidants 2024,13, 395 31 of 31
231.
Wang, D.; Yin, K.; Zhang, Y.; Lu, H.; Hou, L.; Zhao, H.; Xing, M. Fluoride induces neutrophil extracellular traps and aggravates
brain inflammation by disrupting neutrophil calcium homeostasis and causing ferroptosis. Environ. Pollut. 2023,331, 121847.
[CrossRef] [PubMed]
232.
Yao, L.; Sheng, X.; Dong, X.; Zhou, W.; Li, Y.; Ma, X.; Song, Y.; Dai, H.; Du, Y. Neutrophil extracellular traps mediate TLR9/Merlin
axis to resist ferroptosis and promote triple negative breast cancer progression. Apoptosis 2023,28, 1484–1495. [CrossRef]
233.
Zhang, H.; Liu, J.; Zhou, Y.; Qu, M.; Wang, Y.; Guo, K.; Shen, R.; Sun, Z.; Cata, J.P.; Yang, S.; et al. Neutrophil extracellular traps
mediate m(6)A modification and regulates sepsis-associated acute lung injury by activating ferroptosis in alveolar epithelial cells.
Int. J. Biol. Sci. 2022,18, 3337–3357. [CrossRef] [PubMed]
234.
Lv, T.; Xiong, X.; Yan, W.; Liu, M.; Xu, H.; He, Q. Mitochondrial general control of amino acid synthesis 5 like 1 promotes
nonalcoholic steatohepatitis development through ferroptosis-induced formation of neutrophil extracellular traps. Clin. Transl.
Med. 2023,13, e1325. [CrossRef] [PubMed]
235.
Galicia, G.; Boulianne, B.; Pikor, N.; Martin, A.; Gommerman, J.L. Secondary B cell receptor diversification is necessary for T
cell mediated neuro-inflammation during experimental autoimmune encephalomyelitis. PLoS ONE 2013,8, e61478. [CrossRef]
[PubMed]
236.
Hartlehnert, M.; Borsch, A.L.; Li, X.; Burmeister, M.; Gerwien, H.; Schafflick, D.; Heming, M.; Lu, I.N.; Narayanan, V.; Strecker,
J.K.; et al. Bcl6 controls meningeal Th17-B cell interaction in murine neuroinflammation. Proc. Natl. Acad. Sci. USA 2021,118,
e2023174118. [CrossRef] [PubMed]
237.
Mundt, S.; Mrdjen, D.; Utz, S.G.; Greter, M.; Schreiner, B.; Becher, B. Conventional DCs sample and present myelin antigens in the
healthy CNS and allow parenchymal T cell entry to initiate neuroinflammation. Sci. Immunol. 2019,4, eaau8380. [CrossRef]
238.
Rayasam, A.; Kijak, J.A.; Dallmann, M.; Hsu, M.; Zindl, N.; Lindstedt, A.; Steinmetz, L.; Harding, J.S.; Harris, M.G.; Karman,
J.; et al. Regional Distribution of CNS Antigens Differentially Determines T-Cell Mediated Neuroinflammation in a CX3CR1-
Dependent Manner. J. Neurosci. 2018,38, 7058–7071. [CrossRef] [PubMed]
239.
Harrer, C.; Otto, F.; Pilz, G.; Haschke-Becher, E.; Trinka, E.; Hitzl, W.; Wipfler, P.; Harrer, A. The CXCL13/CXCR5-chemokine axis
in neuroinflammation: Evidence of CXCR5+CD4 T cell recruitment to CSF. Fluids Barriers CNS 2021,18, 40. [CrossRef] [PubMed]
240.
Williams, J.L.; Manivasagam, S.; Smith, B.C.; Sim, J.; Vollmer, L.L.; Daniels, B.P.; Russell, J.H.; Klein, R.S. Astrocyte-T cell crosstalk
regulates region-specific neuroinflammation. Glia 2020,68, 1361–1374. [CrossRef] [PubMed]
241.
Luoqian, J.; Yang, W.; Ding, X.; Tuo, Q.Z.; Xiang, Z.; Zheng, Z.; Guo, Y.J.; Li, L.; Guan, P.; Ayton, S.; et al. Ferroptosis promotes
T-cell activation-induced neurodegeneration in multiple sclerosis. Cell Mol. Immunol. 2022,19, 913–924. [CrossRef] [PubMed]
242.
Gross, C.C.; Meyer, C.; Bhatia, U.; Yshii, L.; Kleffner, I.; Bauer, J.; Troscher, A.R.; Schulte-Mecklenbeck, A.; Herich, S.; Schneider-
Hohendorf, T.; et al. CD8(+) T cell-mediated endotheliopathy is a targetable mechanism of neuro-inflammation in Susac syndrome.
Nat. Commun. 2019,10, 5779. [CrossRef] [PubMed]
243.
Liang, J.; Shen, Y.; Wang, Y.; Huang, Y.; Wang, J.; Zhu, Q.; Tong, G.; Yu, K.; Cao, W.; Wang, Q.; et al. Ferroptosis participates in
neuron damage in experimental cerebral malaria and is partially induced by activated CD8(+) T cells. Mol. Brain 2022,15, 57.
[CrossRef] [PubMed]
244.
Laurent, C.; Dorothee, G.; Hunot, S.; Martin, E.; Monnet, Y.; Duchamp, M.; Dong, Y.; Legeron, F.P.; Leboucher, A.; Burnouf, S.;
et al. Hippocampal T cell infiltration promotes neuroinflammation and cognitive decline in a mouse model of tauopathy. Brain
2017,140, 184–200. [CrossRef] [PubMed]
245.
Yshii, L.; Pasciuto, E.; Bielefeld, P.; Mascali, L.; Lemaitre, P.; Marino, M.; Dooley, J.; Kouser, L.; Verschoren, S.; Lagou, V.; et al.
Astrocyte-targeted gene delivery of interleukin 2 specifically increases brain-resident regulatory T cell numbers and protects
against pathological neuroinflammation. Nat. Immunol. 2022,23, 878–891. [CrossRef] [PubMed]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual
author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to
people or property resulting from any ideas, methods, instructions or products referred to in the content.
... Ferroptosis is a recently discovered form of cell death controlled by iron [13]. Prior research has indicated a strong connection between ferroptosis and neuroinflammation [14,15]. Studies have shown that neuroinflammation can trigger intracellular iron buildup, speed up iron overload, facilitate the synthesis of reactive oxygen species (ROS), and further result in ferroptosis [16,17], whereas ferroptosis enhances neuroinflammatory reactions [18,19]. ...
Article
Full-text available
Neuropathic pain, which refers to pain caused by a lesion or disease of the somatosensory system, represents a wide variety of peripheral or central disorders. Treating neuropathic pain is quite demanding, primarily because of its intricate underlying etiological mechanisms. The central nervous system relies on microglia to maintain balance, as they are associated with serving primary immune responses in the brain next to cell communication. Ferroptosis, driven by phospholipid peroxidation and regulated by iron, is a vital mechanism of cell death regulation. Neuroinflammation can be triggered by ferroptosis in microglia, which contributes to the release of inflammatory cytokines. Conversely, neuroinflammation can induce iron accumulation in microglia, resulting in microglial ferroptosis. Accumulating evidence suggests that neuroinflammation, characterized by glial cell activation and the release of inflammatory substances, significantly exacerbates the development of neuropathic pain. By inhibiting microglial ferroptosis, it may be possible to prevent neuroinflammation and subsequently alleviate neuropathic pain. The activation of the homopentameric α7 subtype of the neuronal nicotinic acetylcholine receptor (α7nAChR) has the potential to suppress microglial activation, transitioning M1 microglia to an M2 phenotype, facilitating the release of anti-inflammatory factors, and ultimately reducing neuropathic pain. Recent years have witnessed a growing recognition of the regulatory role of α7nAChR in ferroptosis, which could be a potential target for treating neuropathic pain. This review summarizes the mechanisms related to α7nAChR and the progress of ferroptosis in neuropathic pain according to recent research. Such an exploration will help to elucidate the relationship between α7nAChR, ferroptosis, and neuroinflammation and provide new insights into neuropathic pain management.
Article
Full-text available
Ferroptosis is a non-apoptotic form of regulated cell death characterized by the lethal accumulation of iron-dependent membrane-localized lipid peroxides. It acts as an innate tumor suppressor mechanism and participates in the biological processes of tumors. Intriguingly, mesenchymal and dedifferentiated cancer cells, which are usually resistant to apoptosis and traditional therapies, are exquisitely vulnerable to ferroptosis, further underscoring its potential as a treatment approach for cancers, especially for refractory cancers. However, the impact of ferroptosis on cancer extends beyond its direct cytotoxic effect on tumor cells. Ferroptosis induction not only inhibits cancer but also promotes cancer development due to its potential negative impact on anticancer immunity. Thus, a comprehensive understanding of the role of ferroptosis in cancer is crucial for the successful translation of ferroptosis therapy from the laboratory to clinical applications. In this review, we provide an overview of the recent advancements in understanding ferroptosis in cancer, covering molecular mechanisms, biological functions, regulatory pathways, and interactions with the tumor microenvironment. We also summarize the potential applications of ferroptosis induction in immunotherapy, radiotherapy, and systemic therapy, as well as ferroptosis inhibition for cancer treatment in various conditions. We finally discuss ferroptosis markers, the current challenges and future directions of ferroptosis in the treatment of cancer.
Article
Full-text available
Background: Cingulin (CGN) is a pivotal cytoskeletal adaptor protein located at tight junctions. This study investigates the link between CGN mutation and increased cancer susceptibility through genetic and mechanistic analyses and proposes a potential targeted therapeutic approach. Methods: In a high-cancer-density family without known pathogenic variants, we performed tumor-targeted and germline whole-genome sequencing to identify novel cancer-associated variants. Subsequently, these variants were validated in a 222 cancer patient cohort, and CGN c.3560C > T was identified as a potential cancer-risk allele. Both wild-type (WT) (c.3560C > C) and variant (c.3560C > T) were transfected into cancer cell lines and incorporated into orthotopic xenograft mice model for evaluating their effects on cancer progression. Western blot, immunofluorescence analysis, migration and invasion assays, two-dimensional gel electrophoresis with mass spectrometry, immunoprecipitation assays, and siRNA applications were used to explore the biological consequence of CGN c.3560C > T. Results: In cancer cell lines and orthotopic animal models, CGN c.3560C > T enhanced tumor progression with reduced sensitivity to oxaliplatin compared to the CGN WT. The variant induced downregulation of epithelial marker, upregulation of mesenchymal marker and transcription factor, which converged to initiate epithelial-mesenchymal transition (EMT). Proteomic analysis was conducted to investigate the elements driving EMT in CGN c.3560C > T. This exploration unveiled overexpression of IQGAP1 induced by the variant, contrasting the levels observed in CGN WT. Immunoprecipitation assay confirmed a direct interaction between CGN and IQGAP1. IQGAP1 functions as a regulator of multiple GTPases, particularly the Rho family. This overexpressed IQGAP1 was consistently associated with the activation of Rac1, as evidenced by the analysis of the cancer cell line and clinical sample harboring CGN c.3560C > T. Notably, activated Rac1 was suppressed following the downregulation of IQGAP1 by siRNA. Treatment with NSC23766, a selective inhibitor for Rac1-GEF interaction, resulted in the inactivation of Rac1. This intervention mitigated the EMT program in cancer cells carrying CGN c.3560C > T. Consistently, xenograft tumors with WT CGN showed no sensitivity to NSC23766 treatment, but NSC23766 demonstrated the capacity to attenuate tumor growth harboring c.3560C > T. Conclusions: CGN c.3560C > T leads to IQGAP1 overexpression, subsequently triggering Rac1-dependent EMT. Targeting activated Rac1 is a strategy to impede the advancement of cancers carrying this specific variant.
Article
Full-text available
Background Inflammation plays a pivotal role in the pathogenesis of Parkinson’s disease (PD). However, the correlation between peripheral inflammatory markers and the severity of PD remains unclear. Methods The following items in plasma were collected for assessment among patients with PD (n = 303) and healthy controls (HCs; n = 303) were assessed for the neutrophil-to-lymphocyte ratio (NLR), lymphocyte-to-monocyte ratio (LMR) and neutrophil-to-high-density-lipoprotein ratio (NHR) in plasma, and neuropsychological assessments were performed for all patients with PD. Spearman rank or Pearson correlation was used to evaluate the correlation between the NLR, the LMR and the NHR and the severity of PD. Receiver operating characteristic (ROC) curves were used to evaluate the diagnostic performance of the NLR, LMR and NHR for PD. Results The plasma NLR and NHR were substantially higher in patients with PD than in HCs, while the plasma LMR was substantially lower. The plasma NLR was positively correlated with Hoehn and Yahr staging scale (H&Y), Unified Parkinson’s Disease Rating Scale (UPDRS), UPDRS-I, UPDRS-II, and UPDRS-III scores. Conversely, it exhibited a negative relationship with Mini-Mental State Examination (MMSE) and Montreal Cognitive Assessment (MoCA) scores. Furthermore, the plasma NHR was positively correlated with H&Y, UPDRS, UPDRS-I, UPDRS-II and UPDRS-III scores. Moreover, negative associations were established between the plasma LMR and H&Y, UPDRS, UPDRS-I, UPDRS-II, and UPDRS-III scores. Finally, based on the ROC curve analysis, the NLR, LMR and NHR exhibited respectable PD discriminating power. Conclusion Our research indicates that a higher NLR and NHR and a lower LMR may be relevant for assessing the severity of PD and appear to be promising disease-state biomarker candidates.
Article
Full-text available
Background Microglia-mediated neuroinflammation in Alzheimer’s disease (AD) is not only a response to pathophysiological events, but also plays a causative role in neurodegeneration. Cytoplasmic cysteinyl-tRNA synthetase (CARS) is considered to be a stimulant for immune responses to diseases; however, it remains unknown whether CARS is involved in the pathogenesis of AD. Methods Postmortem human temporal cortical tissues at different Braak stages and AD patient-derived serum samples were used to investigate the changes of CARS levels in AD by immunocytochemical staining, real-time PCR, western blotting and ELISA. After that, C57BL/6J and APP/PS1 transgenic mice and BV-2 cell line were used to explore the role of CARS protein in memory and neuroinflammation, as well as the underlying mechanisms. Finally, the associations of morphological features among CARS protein, microglia and dense-core plaques were examined by immunocytochemical staining. Results A positive correlation was found between aging and the intensity of CARS immunoreactivity in the temporal cortex. Both protein and mRNA levels of CARS were increased in the temporal cortex of AD patients. Immunocytochemical staining revealed increased CARS immunoreactivity in neurons of the temporal cortex in AD patients. Moreover, overexpression of CARS in hippocampal neurons induced and aggravated cognitive dysfunction in C57BL/6J and APP/PS1 mice, respectively, accompanied by activation of microglia and the TLR2/MyD88 signaling pathway as well as upregulation of proinflammatory cytokines. In vitro experiments showed that CARS treatment facilitated the production of proinflammatory cytokines and the activation of the TLR2/MyD88 signaling pathway of BV-2 cells. The accumulation of CARS protein occurred within dense-core Aβ plaques accompanied by recruitment of ameboid microglia. Significant upregulation of TLR2/MyD88 proteins was also observed in the temporal cortex of AD. Conclusions The findings suggest that the neuronal CARS drives neuroinflammation and induces memory deficits, which might be involved in the pathogenesis of AD.
Article
Mammalian physiology and cellular function are subject to significant oscillations over the course of every 24-hour day. It is likely that these daily rhythms will affect function as well as mechanisms of disease in the central nervous system. In this review, we attempt to survey and synthesize emerging studies that investigate how circadian biology may influence the neurovascular unit. We examine how circadian clocks may operate in neural, glial, and vascular compartments, review how circadian mechanisms regulate cell-cell signaling, assess interactions with aging and vascular comorbidities, and finally ask whether and how circadian effects and disruptions in rhythms may influence the risk and progression of pathophysiology in cerebrovascular disease. Overcoming identified challenges and leveraging opportunities for future research might support the development of novel circadian-based treatments for stroke.