ArticlePDF AvailableLiterature Review

Post-ischemic inflammation regulates neural damage and protection

Authors:

Abstract and Figures

Post-ischemic inflammation is important in ischemic stroke pathology. However, details of the inflammation process, its resolution after stroke and its effect on pathology and neural damage have not been clarified. Brain swelling, which is often fatal in ischemic stroke patients, occurs at an early stage of stroke due to endothelial cell injury and severe inflammation by infiltrated mononuclear cells including macrophages, neutrophils, and lymphocytes. At early stage of inflammation, macrophages are activated by molecules released from necrotic cells [danger-associated molecular patterns (DAMPs)], and inflammatory cytokines and mediators that increase ischemic brain damage by disruption of the blood-brain barrier are released. After post-ischemic inflammation, macrophages function as scavengers of necrotic cell and brain tissue debris. Such macrophages are also involved in tissue repair and neural cell regeneration by producing tropic factors. The mechanisms of inflammation resolution and conversion of inflammation to neuroprotection are largely unknown. In this review, we summarize information accumulated recently about DAMP-induced inflammation and the neuroprotective effects of inflammatory cells, and discuss next generation strategies to treat ischemic stroke.
Content may be subject to copyright.
REVIEW ARTICLE
published: 14 October 2014
doi: 10.3389/fncel.2014.00319
Post-ischemic inflammation regulates neural damage and
protection
Takashi Shichita1,2 , Minako Ito1and Akihiko Yoshimura1*
1Department of Microbiology and Immunology, School of Medicine, Keio University,Tokyo, Japan
2Precursory Research for Embryonic Science andTechnology, Japan Science and Technology Agency, Tokyo, Japan
Edited by:
Arthur Liesz, University Hospital
Munich, Germany
Reviewed by:
Mathias Gelderblom, University
Medical Center Hamburg-Eppendorf,
Germany
Alexander Dressel, University
Medicine Greifswald, Germany
*Correspondence:
Akihiko Yoshimura, Department of
Microbiology and Immunology, School
of Medicine, Keio University, 35
Shinanomachi, Shinjuku-ku, Tokyo
160-8582, Japan
e-mail: yoshimura@a6.keio.jp
Post-ischemic inflammation is important in ischemic stroke pathology. However, details
of the inflammation process, its resolution after stroke and its effect on pathology and
neural damage have not been clarified. Brain swelling, which is often fatal in ischemic
stroke patients, occurs at an early stage of stroke due to endothelial cell injury and severe
inflammation by infiltrated mononuclear cells including macrophages, neutrophils, and
lymphocytes. At early stage of inflammation, macrophages are activated by molecules
released from necrotic cells [danger-associated molecular patterns (DAMPs)], and inflam-
matory cytokines and mediators that increase ischemic brain damage by disruption of the
blood–brain barrier are released. After post-ischemic inflammation, macrophages function
as scavengers of necrotic cell and brain tissue debris. Such macrophages are also involved
in tissue repair and neural cell regeneration by producing tropic factors. The mechanisms
of inflammation resolution and conversion of inflammation to neuroprotection are largely
unknown. In this review, we summarize information accumulated recently about DAMP-
induced inflammation and the neuroprotective effects of inflammatory cells, and discuss
next generation strategies to treat ischemic stroke.
Keywords: damage-associated molecular patterns (DAMPs), inflammation, cytokines, inflammasome, resolution
of inflammation
INTRODUCTION
Inflammation is implicated in almost all of central nervous sys-
tem (CNS) diseases (Lo, 2010;Moskowitz et al., 2010;Iadecola
and Anrather, 2011). Neurodegeneration, infection, trauma, and
ischemia stimulate immune responses in the brain, although to
varying degrees. The process in neuronal injury involves various
intracellular mechanisms (abnormal metabolism and degenera-
tion of protein, dysfunction of organelles, etc.), which cause the
activation of microglia and the infiltration of circulating immune
cells (Lo, 2010). Inflammation may not be always main process in
the pathology of CNS diseases; nonetheless, the distinct charac-
teristics of ischemic stroke are large amount of necrotic neuronal
death and extreme infiltration of immune cells (Moskowitz et al.,
2010;Iadecola and Anrather, 2011).
Severe inflammation causes cerebral swelling, which is often
fatal in ischemic stroke patients. Broad necrotic lesion gener-
ates abundant inflammatory mediators and damage-associated
molecular patterns (DAMPs), which enhance the chemotaxis of
circulating immune cells and make them more efficient partici-
pants to promote inflammation (Moskowitz et al., 2010;Iadecola
and Anrather, 2011). Cerebral inflammation exaggerates vascular
dysfunction and induces further neuronal cell death (Dirnagl et al.,
1999). Thus, post-ischemic inflammation is an essential process in
the pathophysiology of ischemic stroke and is closely related to the
prognosis after stroke (Dirnagl et al., 1999;Lo, 2010;Moskowitz
et al., 2010;Iadecola and Anrather, 2011). In addition, inflamma-
tion is generally considered useful for the clearance of the large
amount of debris caused by brain cell necrotic death (Moskowitz
et al., 2010;Iadecola and Anrather, 2011). Inflammation, resolu-
tion of inflammation, and repair of neural damage are sequential
pivotal events after stroke. To clarify the detailed mechanisms of
each step of cerebral inflammation is indispensable to develop next
generation therapies for ischemic stroke. The molecular basis of
these steps is now being clarified by the recent accumulating evi-
dences. We summarize these findings and discuss the principles of
post-ischemic inflammation from beginning to end.
INFLAMMATORY DAMPs
Brain ischemia induces various large metabolic changes in brain
cells. Hypoxic stress, nutrients stress, and ER stress will cause cell
death and trigger post-ischemic inflammation. Although receptors
for pathogens such as Toll-like receptors (TLRs) are thought to be
involved in early step of inflammation, brain is a sterile organ.
Thus, endogenous molecules, i.e., DAMPs derived from injured
brain cells, must trigger the inflammatory response in immune
cells (Ta bl e 1 ). These DAMPs induce the activation of TLRs and
other pattern recognition receptors [receptor for advanced glyca-
tion end products (RAGE) and c-type lectin receptors], which
promote inflammatory mediator expression and tissue injury
(Figure 1;Tang et al., 2007;Yanai et al., 2009;Suzuki et al., 2013).
Recent scientific advances have suggested the existence of various
types of DAMPs in ischemic brain.
NUCLEIC ACIDS AND NUCLEOTIDES
Various intracellular components are released into the
extracellular space by necrotic brain cell death. Among these,
Frontiers in Cellular Neuroscience www.frontiersin.org October 2014 |Volume 8 |Article 319 |1
Shichita et al. Inflammatory mechanisms after ischemic stroke
Table 1 |List of inflammatory DAMPs.
DAMPs Receptor Reference
Signal 1 Nucleic acid Mitochondrial DNA TLR9 Zhang et al. (2010),Sun etal. (2013),Maeda and Fadeel (2014),Walko et al.
(2014),Wenceslau etal. (2014)
Self RNA, DNA TLR7,9 Hyakkoku etal. (2010),Kawai and Akira (2010),Brea et al. (2011),Stevens et al.
(2011),Leung et al. (2012)
Lipid Carboxyalkylpyrroles TLR2 West et al. (2010)
Oxidized phospholipids CD36 Cho et al. (2005),Gao et al. (2006),Abe et al. (2010),Haider etal. (2011),Miller
et al. (2011),Ho et al. (2012),Matt et al. (2013)
Protein HMGB1 TLR2,4, RAGE Qiu et al. (2008),Zhang et al. (2011)
Peroxiredoxin TLR2,4 Shichita et al. (2012),Kuang et al. (2014)
S100A8, A9 TLR4 Tsai et al. (2014)
Mrp8, 14 TLR4 Loser et al. (2010)
CIRP TLR2,4 Qiang et al. (2013)
Signal 2 Nucleotide ATP P2X, P2Y Martinon et al. (2002),Abulafia et al. (2009),Ceruti et al. (2009),Denes et al.
(2013),Fann et al. (2013),Yang etal. (2014),
Lipid Phospholipids ? Clemens et al. (1996),Bonventre etal. (1997),Muralikrishna Adibhatla and
Hatcher (2006),Shanta et al. (2012),Iyer etal. (2013),Zhong etal. (2013)
Protein ASC specks ? Baroja-Mazo et al. (2014),Franklin et al. (2014)
FIGURE 1 |Mechanisms of post-ischemic inflammation. DAMPs are
released into extracellular compartment and activate infiltrating immune
cells by two ways: Signal 1 (via the activation of pattern recognition
receptor) and Signal 2 (via the activation of inflammasome). Various
inflammatory cytokines promote neuronal injury, and induce further
inflammation mediated by T cells in subacute phase. After days and week
after stroke onset, the resolution of post-ischemic inflammation is brought
by the clearance of debris including DAMPs or inflammatory mediators,
and the production of anti-inflammatory molecules or neurotrophic factors.
In this recover phase, inflammatory immune cells turn into neuroprotective
cells.
nucleic acids and nucleotides are major DAMPs that have recently
received much attention. Mitochondrial DNA released by cellu-
lar injury can be detected as DAMPs by immune cells, because
mitochondria are considered to have a symbiotic origin that
carries numerous characteristics resembling bacteria. Mitochon-
drial DNA is a sensor molecule of innate immunity by activating
TLR9 and can be detected in cerebrospinal fluid after traumatic
brain injury (Zhang et al., 2010;Walko et al., 2014). Vascular per-
meability is also increased by circulating mitochondrial DNA
after injury (Sun et al., 2013;Wenceslau et al., 2014). Recently
accumulated data indicates that mitochondrial DAMPs could
be an important candidate for the trigger of post-ischemic
inflammation, even if there is not yet any direct evidence
(Maeda and Fadeel, 2014).
Self RNA and DNA (in complex with LL37 peptide) acti-
vate immune cells via TLR7 or TLR9 (Kawai and Akira, 2010).
TLR7 is associated with the deterioration in ischemic stroke
patients; in contrast, ischemic brain damage was not reduced
in TLR9-deficient mice (Hyakkoku et al., 2010;Brea et al., 2011).
Several reports demonstrate the implications of TLR7 and TLR9
in ischemic preconditioning. In these articles, the pretreatment
using a TLR7 or TLR9 agonist reveals significant neuroprotection
after cerebral ischemia by activating interferon regulatory factor
3/7- (IRF3/7)-induced type I interferon (IFN) signaling pathway
(Stevens et al., 2011;Leung et al., 2012). Although the interaction
between self nucleic acids and TLRs in the ischemic brain remains
controversial, ischemic preconditioning via the TLR7 or TLR9
signaling pathway may represent a therapeutic strategy.
Purines (ATP and UTP) released from injured brain cells and
their receptors, P2X and P2Y, function as alerting signals in CNS
(Ceruti et al., 2009). Importantly, ATP also activates inflamma-
somes, which are large multimolecular complexes that control the
activity of the proteolytic enzyme caspase-1 that cleaves pro-IL-1β
to an active 17 kDa form (Martinon et al., 2002). The activation of
Frontiers in Cellular Neuroscience www.frontiersin.org October 2014 |Volume 8 |Article 319 |2
Shichita et al. Inflammatory mechanisms after ischemic stroke
the NLRP1 or NALP3 inflammasome has been recently reported to
promote post-ischemic inflammation and neuronal death (Abu-
lafia et al., 2009;Fann et al., 2013;Yang et al., 2014). Because IL-1β
produced from both infiltrating immune cells and brain cells is
important (Denes et al., 2013), it should be clarified how the
inflammasome is activated in ischemic brain or hematopoietic
cells. Inhibition of the inflammasome activation pathway may be
a possible therapeutic strategy for ischemic stroke.
LIPIDS
Various types of lipids are also important regulators of innate
immunity. For example, oxidized low density lipoprotein (oxLDL)
is a popular inflammatory mediator, which activates TLRs
through binding with its receptor, CD36 (Stewart et al., 2010).
Although the function of oxLDL in ischemic brain remains
unclear, recent research has indicated that end products of
lipid oxidation may be implicated in cerebral post-ischemic
inflammation (Uchida, 2013). Carboxyalkylpyrroles, which are
generated in inflammatory tissue, activate TLR2 and promote
angiogenesis in ischemic organs (West et al., 2010). Oxidized
phospholipids are also generated during cerebral inflammation
and are considered to be DAMPs (Gao et al., 2006;Haider et al.,
2011;Ho et al., 2012). Oxidized phospholipids are CD36 lig-
ands that promote inflammation via TLR2 activation in ischemic
brain (Cho et al., 2005;Abe et al., 2010). The recognition and
endocytosis of oxidized lipids by pattern recognition receptors
could regulate post-ischemic inflammation (Miller et al., 2011;
Matt et al., 2013).
Phospholipids could also be inflammasome activators. Phos-
pholipid metabolism is drastically altered by cerebral ischemia
(Shanta et al., 2012). There are several reports showing the acti-
vation of phospholipase A2 (PLA2) in ischemic brain, which
results in hydrolysis of membrane phospholipids (Clemens
et al., 1996;Bonventre et al., 1997;Muralikrishna Adibhatla
and Hatcher, 2006). Phospholipid hydrolysis and mitochon-
drial dysfunction induced by cerebral ischemia generate reac-
tive oxygen species (ROS). Two recent studies have identi-
fied both ROS-dependent and ROS-independent pathways for
inflammasome activation. The former is demonstrated by a
charged phospholipid liposome that consecutively induces ROS-
dependent calcium influx and NLRP3 inflammasome activation
(Zhong et al., 2013). In the latter case, mitochondrial cardi-
olipin has been reported to directly bind to and activate the
NLRP3 inflammasome (Iyer et al., 2013). Thus, the metabolism
and modification of lipids during cerebral ischemia may be
closely associated with the post-ischemic inflammation start
signal.
PROTEINS
High mobility group box 1 (HMGB1) and peroxiredoxin (Prx)
family proteins are two major DAMPs in ischemic brain. There is
a difference in the functional phase of these two proteins (Shi-
chita et al., 2012). HMGB1, which is included in the nucleus
of brain cells, is released extracellularly at the hyperacute phase
(several hours after the stroke onset; Qiu et al., 2008). On the
other hand, Prx family proteins function at the acute and suba-
cute phases (12–72 h after the onset), especially in the penumbral
area (Shichita et al., 2012). This is because Prx family pro-
tein expression is induced by an intracellular increase in ROS,
which results from ischemic change. HMGB1 directly breaks
down the blood–brain barrier and increases vascular permeability
(Zhang et al., 2011). However, Prx directly induces the activa-
tion of infiltrating immune cells via TLR signaling. Ligustilide has
been reported as a therapeutic candidate that suppresses cerebral
post-ischemic inflammation by inhibiting the Prx/TLR4 signaling
pathway (Kuang et al., 2014).
S100A8, S100A9, Mrp8, Mrp14, and cold-inducible RNA bind-
ing protein (CIRP) have also been reported to be protein DAMPs,
although their relevance in post-ischemic inflammation has not
yet been clarified (Loser et al., 2010;Qiang et al., 2013;Tsai et al.,
2014). Inflammatory responses by these protein DAMPs occur
through the activation of TLR2, TLR4, and RAGE. TLR2 and TLR4
signaling pathways are essential for sterile inflammation, includ-
ing ischemic stroke (Chen et al., 2007). TLR2-blocking antibody
is neuroprotective against ischemic brain injury (Ziegler et al.,
2011). Similarly, resatorvid, which inhibits the TLR4 signaling
pathway, attenuates ischemic brain injury and also suppresses
Nox4-induced oxidative stress and neuronal apoptosis (Suzuki
et al., 2012). It is also possible that DAMP-mediated TLR activation
requires other adaptor molecules (Chun and Seong, 2010). CD14,
a TLR4 co-receptor, may be implicated in post-ischemic inflam-
mation (Reed-Geaghan et al., 2009). Heat shock protein gp96 is
another candidate molecule that functions as an adaptor for both
TLR2 and TLR4 (Yang et al., 2007).
It is not known whether protein DAMPs can activate inflamma-
somes. Recently, aggregated ASC (apoptosis-associated speck-like
protein containing a caspase recruitment domain) has been
reported to be released into the extracellular space after cell death
and it activates inflammasomes in the surrounding immune cells
(Baroja-Mazo et al., 2014;Franklin et al., 2014). Inflammasome
activation, which occurs through ASC polymerization, results
in caspase-1 activation and pyroptotic cell death. Extracellularly
released ASC is internalized by surrounding macrophages and
induces lysosomal damage and inflammasome activation. These
mechanisms of inflammasome activation remain to be elucidated
in ischemic brain injury.
OTHER INFLAMMATORY DAMPs
Basic research may neglect the influence of aging and life habits by
using healthy young rodents. These are important factors for the
generation of DAMPs. Aging and continuous high serum glucose
levels increase lipid peroxidation andAGEs in body systems (Basta
et al., 2004;Cai et al., 2014). AGEs are proteins that are modified
by sugar, through the Amadori and Maillard reaction. AGEs are
found in chronic lesions; for example, the amyloid deposits that
are surrounded by macrophages in patients with dialysis-related
amyloidosis (Miyata et al., 1993). Thus, AGEs usually take a long
time (more than a month) to generate; however, AGEs can be gen-
erated in a short period of time during inflammation (Weil, 2012).
Glyoxal and glyceraldehyde induce AGE formation within 1 week
(Takeuchi et al., 2001). In addition, the pivotal role of the RAGE
in post-ischemic inflammation has been demonstrated (Muham-
mad et al., 2008). AGEs can be a potential DAMP, especially in aged
human ischemic stroke patients.
Frontiers in Cellular Neuroscience www.frontiersin.org October 2014 |Volume 8 |Article 319 |3
Shichita et al. Inflammatory mechanisms after ischemic stroke
INFLAMMATION SUPPRESSION AND RESOLUTION
Activated immune cells and brain cells are the major play-
ers after various DAMPs trigger post-ischemic inflammation.
These cells produce inflammatory cytokines, chemokines, and
other cytotoxic mediators, and this leads to prolonged inflam-
mation and progressive brain edema during several days after
the stroke onset (Figure 1). However, post-ischemic inflam-
mation rarely lasts for a long period of time, and the most
intense inflammatory phase takes place within 7 days after stroke
onset (Dirnagl et al., 1999;Iadecola and Anrather, 2011). In
this phase, the number of infiltrating immune cells decreases
remarkably, and remaining immune cells in ischemic brain
produce anti-inflammatory or neurotrophic factors (Shichita
et al., 2009;Smirkin et al., 2010). For example, the detailed
mechanisms about the infiltration and the change to anti-
inflammatory phenotype of neutrophils have been recently clar-
ified (Cuartero et al., 2013;Gorina et al., 2014). The period
of cerebral post-ischemic inflammation always ends, and thus,
the mechanisms of its resolution must exist in ischemic
brain.
Three major points on the resolution of inflammation have
been discussed in a recent publication (Buckley et al., 2012).
These points are the production of anti-inflammatory media-
tors, the depletion of inflammatory mediators, and the induction
of anti-inflammatory immune cells. After post-ischemic inflam-
mation, infiltrating macrophages turn into anti-inflammatory
macrophages, which produce neurotropic factors and clear
necrotic debris. Inflammatory DAMPs will also be implicated
in the induction of anti-inflammatory macrophages, although
its mechanism still remains to be clarified. We introduce recent
advantages of the relationship between post-ischemic inflamma-
tion and its resolution.
ANTI-INFLAMMATORY MEDIATOR PRODUCTION
Many molecules have been reported to be neuroprotective factors.
However, most of these molecules failed to improve neurological
deficits in ischemic stroke patients, even if they are effective in
animal stroke models. It has been suggested that neuroprotection
alone is not sufficient to improve the prognosis of human ischemic
stroke. Anti-inflammatory mechanisms in the entire brain and
how these mechanisms are triggered needs to be determined.
Because most brain cells are dead in the ischemic region several
days after the stroke onset, infiltrating immune cells and reactive
glial cells could be major players in the tissue repair. In practice,
accelerating their effect is a potential next generation therapeutic
strategy, and direct in vivo reprogramming of reactive glial cells
into functional neurons after cerebral injury by retroviral trans-
duction of the NeuroD1 gene was recently reported (Guo et al.,
2014).
IL-10 and TGF-βare major anti-inflammatory molecules
in various organ injuries. Both are produced by infiltrating
immune cells and reactive glial cells after ischemic brain injury.
Viral overexpression of IL-10 in ischemic brain is neuroprotec-
tive (Ooboshi et al., 2006). One recent report demonstrated the
anti-inflammatory effect of TGF-βby inhibiting excessive neu-
roinflammation during the subacute phase of brain ischemia
(Cekanaviciute et al., 2014). Although the anti-inflammatory
effects of IL-10 and TGF-βhave been pivotal, it remains to be
clarified whether these effects last up to 1 week after the stroke
onset (Pál et al., 2012). If the mechanisms for stimulating TGF-β
and IL-10 expression can be controlled, this may become a strong
therapeutic method.
DEPLETION OF INFLAMMATORY MEDIATORS AND CELLS
Infiltrating immune cells and reactive glial cells produce various
inflammatory mediators. TNF-αand IL-1βdirectly induce neu-
ronal cell death. IL-23 and IL-1βactivate T cell-mediated innate
immunity and promote secondary ischemic damage during the
subacute phase of ischemic brain injury (Shichita et al., 2009;
Konoeda et al., 2010). The existence of these inflammatory medi-
ators, including DAMPs, prolongs post-ischemic inflammation
and will be a threat to neuronal survival and repair. How-
ever, inflammatory mediator degradation mechanisms remain
mostly unknown. Inflammatory molecules may be degraded by
some enzymes or consumed by receptor-mediated endocytosis.
It is expected that nucleotides and lipids are rapidly metabo-
lized in the ischemic brain, and transfer by the blood stream
or cerebrospinal fluid (CSF) will help to scavenge inflammatory
mediator. Further research should clarify the detailed mech-
anisms to scavenge inflammatory molecules produced in the
ischemic brain, and antibody therapy will be a pivotal thera-
peutic method targeting this potential mechanism. TNF-αand
IL-23 neutralizing antibody have been used clinically for rheuma-
toid arthritis and psoriasis patients, respectively. Natalizumab
is the neutralizing antibody for integrin-α4, which is neces-
sary for T cell infiltration into the inflammatory tissue, and has
already been used for multiple sclerosis (Yednock et al., 1992).
T cell depletion from ischemic brain has received attention as
a potential next generation therapy for ischemic stroke (Meisel
and Meisel, 2011;Wei et al., 2011). Thus, antibody therapies
may be used to help treat ischemic stroke patients in the near
future.
Activation of inflammasomes in immune cells induces the pro-
duction of IL-1βin its mature form, and finally results in the rapid
cell death of the same cells, which is called pyroptosis. Pyroptosis
may be a possible mechanism for the clearance of inflammatory
immune cells. This is supported by the fact that dying cells detected
using the TdT-mediated dUTP nick end labeling (TUNEL) stain-
ing method include macrophages and glial cells in the ischemic
brain (Mabuchi et al., 2000).
INDUCTION OF IMMUNE CELL REPAIR
The repair process for damaged brain tissues and regeneration
of neural cells takes place during resolution of inflammation. It
is difficult to separate this process from the anti-inflammatory
mechanism, because they may overlap each other. We will fur-
ther discuss neuroprotective factors and repairing the damage to
immune cells.
NEUROPROTECTIVE FACTORS
Various growth factors are also produced by immune cells and
glial cells (Gudi et al., 2011). Among these, IGF-1 and FGF-2
are produced by infiltrating macrophages and microglia dur-
ing the recovery phase of ischemic brain injury (which occurs
Frontiers in Cellular Neuroscience www.frontiersin.org October 2014 |Volume 8 |Article 319 |4
Shichita et al. Inflammatory mechanisms after ischemic stroke
1 week after the stroke onset). IGF-1 and FGF-2 improve the
neurological outcome by saving neuron and glial cells from
cell death (Ness et al., 2004;Ikeda et al., 2005;Zhu et al., 2008;
Hill et al., 2012;Lalancette-Hébert et al., 2012). IGF-1 also
enhances repair after ischemic stroke by promoting neural regen-
eration, remyelination, and synaptogenesis (Lecker et al., 2007;
Zhu et al., 2008,2009;Kooijman et al., 2009). Further inves-
tigation should clarify the mechanisms of IGF-1 and FGF-2
induction in the ischemic brain. Recently, transfer of mesenchy-
mal stem cells (MSCs) has been explored as a next generation
therapy for ischemic stroke (Kalladka and Muir, 2014). MSCs
produce various growth factors and promote neuronal survival
and neurogenesis (Calió et al., 2014). By improving the transfer
method, cell therapy may become a pivotal therapeutic strategy
(Guo et al., 2013).
The neuroprotective effect of prostaglandin E2 (PGE2) and
its receptor signaling pathway has received recent attention.
PGE2 has an effect via four distinct G protein-coupled EP
receptors (E-prostanoid: EP1, EP2, EP3, and EP4). The activa-
tion of EP2 signaling has a neuroprotective effect in ischemic
brain injury, which was shown in the significant increase in
infarct volume in mice lacking the EP2 receptor (McCullough
et al., 2004). Similarly, signaling via the EP4 receptor, which is
expressed in both neurons and ischemic endothelial cells, has
neuroprotective effects against ischemic brain injury (Liang et al.,
2011). The administration of an EP4 agonist reduces infarct vol-
ume and neurological deficits. In a neonatal hypoxic-ischemic
encephalopathy model, the inhibition of EP1 receptor signal-
ing or the activation of EP2, EP3, and EP4 receptor signaling
reveals attenuation of the ischemic injury (Taniguchi et al., 2011).
PGE2 and EP receptor signaling pathways have various func-
tions, which are dependent on distinct pathology of cerebral
diseases (Furuyashiki and Narumiya, 2011). Targeting specific
EP receptors in ischemic brain may become a novel therapeutic
method.
Similar to PGE2, some lipids have been reported to have anti-
inflammatory effects and promote neuroregeneration (Serhan,
2014). Cerebral ischemia increases PLA2 activity, which results in
the hydrolysis of phospholipids in the cellular membrane (Shanta
et al., 2012). Although the PLA2 effect itself is cytotoxic because
it disrupts the cellular membrane, PLA2 also generates docosa-
hexaenoic acid (DHA) derivatives and lysophospholipids through
phospholipid hydrolysis (Bonventre et al., 1997). Resolvin and
Neuroprotectin have been investigated for their anti-inflammatory
function in ischemic stroke (Marcheselli et al., 2003;Bazan, 2009).
Lysophospholipids also increase in ischemic brain and promote
neurite outgrowth (Ikeno et al., 2005;Spohr et al., 2011;Shanta
et al., 2012). Regulating the effect of these lipids is expected for the
resolution of post-ischemic inflammation.
NEUROPROTECTIVE CELLS
Inflammatory DAMPs activate glial cells and infiltrating immune
cells to promote post-ischemic inflammation. Paradoxically, this
mechanism results in the infiltrating macrophage cell death and
also induces anti-inflammatory and tissue-repairing immune cells.
Immune cell activation also induces anti-inflammatory cells.
These cells have been called M2 macrophages, in contrast to the
inflammatory M1 macrophages. Many researches have described
the M2 macrophage markers; these markers include: arginase-
1 (Arg1), chitinase3l3 (Ym), and Relmα(Fizz1). These markers
are intracellular enzymes that are implicated in collagen synthesis
and cell division; therefore, M2 enzymes are considered to pro-
mote tissue repair. Arg1 is the only marker that was reported
to function as a neuroprotective enzyme (Estévez et al., 2006).
However, these M2 markers may not be a good indicator for
recovery after ischemic stroke. M2 markers are rapidly expressed
in macrophages by TLR activation or other pattern recognition
receptors, which also induce inflammatory cytokine expression
(Hu et al., 2012). M2 markers appear in ischemic brain mostly
during the same phase as the inflammatory mediators, the M1
markers, are expressed. In addition, the transfer of M2 marker
positive-macrophages has not been reported to be sufficiently
neuroprotective (Desestret et al., 2013).
During post-ischemic inflammation, some populations of
macrophages and microglia become neuroprotective (Lalancette-
Hébert et al., 2007). Galectin-1 has been suggested to be
an inducer of anti-inflammatory macrophage/microglial cells
(Starossom et al., 2012;Quintá et al., 2014). Galectin-1 is pro-
duced by astrocytes and has a neuroprotective effect against
ischemic brain damage (Qu et al., 2011). Thus, the resolu-
tion of post-ischemic inflammation can be enhanced by the
induction of a specific macrophage/microglial cell population.
However, it is not clear whether the M2 markers truly reflect
the neuroprotective function of macrophages and microglial
cells. Suppressing inflammation alone is not enough to pro-
tect the brain from ischemic injury. IGF-1 and FGF-2 pro-
duction seems to be a good index of repairing function
(Lalancette-Hébert et al., 2012).
Further study is required to clarify whether sufficient clear-
ance of inflammatory mediators (including DAMPs) begins
neuronal regeneration after ischemic stroke. A recent study
has suggested that there is a relationship between TLR activa-
tion and neuronal repair (Bohacek et al., 2012). It is possible
that DAMPs triggers the secondary signals, which lead to res-
olution of post-ischemic inflammation, even if the primary
signals via pattern recognition receptors promote ischemic dam-
age. What is this mechanism? The role of immune cells, other
than macrophages and microglia in part of the repair process,
is not fully understood. This understanding may be critical
for the establishment of next generation therapies for ischemic
stroke.
CONCLUSION
Immunity and various physiological mechanisms are implicated in
the triggering, persistence, and resolution of post-ischemic inflam-
mation. Recent accumulating evidences clarify the complexity of
these mechanisms to understand the entire mechanisms. They will
show promising potential targets to develop therapies for ischemic
stroke.
REFERENCES
Abe, T., Shimamura, M., Jackman, K., Kurinami, H., Anrather, J., Zhou, P., etal.
(2010). Key role of CD36 in Toll-like receptor 2 signaling in cerebral ischemia.
Stroke 41, 898–904. doi: 10.1161/STROKEAHA.109.572552
Frontiers in Cellular Neuroscience www.frontiersin.org October 2014 |Volume 8 |Article 319 |5
Shichita et al. Inflammatory mechanisms after ischemic stroke
Abulafia, D. P., de Rivero Vaccari, J. P., Lozano, J. D., Lotocki, G., Keane, R. W.,
and Dietrich, W. D. (2009). Inhibition of the inflammasome complex reduces the
inflammatory response after thromboembolic stroke in mice. J. Cereb. Blood Flow
Metab. 29, 534–544. doi: 10.1038/jcbfm.2008.143
Baroja-Mazo, A., Martín-Sánchez, F., Gomez, A., Martínez, C. M., Amores-Iniesta,
J., Compan,V., et al. (2014). The NLRP3 inflammasome is released as a particulate
danger signal that amplifies the inflammatory response. Nat. Immunol. 15, 738–
748. doi: 10.1038/ni.2919
Basta, G., Schmidt, A. M., and De Caterina, R. (2004). Advanced glycation end
products and vascular inflammation: implications for accelerated atherosclerosis
in diabetes. Cardiovasc. Res. 63, 582–592. doi: 10.1016/j.cardiores.2004.05.001
Bazan, N. G. (2009). Neuroprotectin D1-mediated anti-inflammatory and survival
signaling in stroke, retinal degenerations, and Alzheimer’s disease. J. Lipid Res.
50, S400–S405. doi: 10.1194/jlr.R800068-JLR200
Bohacek, I., Cordeau, P., Lalancette-Hébert, M., Gorup, D., Weng, Y. C.,
Gajovic, S., et al. (2012). Toll-like receptor 2 deficiency leads to delayed exac-
erbation of ischemic injury. J. Neuroinflammation 9, 191. doi: 10.1186/1742-
2094-9-191
Bonventre, J. V., Huang, Z., Taheri, M. R., O’Leary, E., Li, E., Moskowitz, M. A.,
et al. (1997). Reduced fertility and postischaemic brain injury in mice deficient
in cytosolic phospholipase A2. Nature 390, 622–625. doi: 10.1038/37635
Brea, D., Sobrino, T., Rodríguez-Yáñez, M., Ramos-Cabrer, P., Agulla, J., Rodríguez-
González, R., et al. (2011). Toll-like receptors 7 and 8 expression is associated with
poor outcome and greater inflammatory response in acute ischemic stroke. Clin.
Imunol. 139, 193–198. doi: 10.1016/j.clim.2011.02.001
Buckley, C. D.,Gilroy, D. W., Serhan, C. N., Stockinger, B., and Tak, P. P. (2012). The
resolution of inflammation. Nat. Rev. Immunol. 13, 59–66. doi: 10.1038/nri3362
Cai, W., Uribarri, J., Zhu, L., Chen, X., Swamy, S., Zhao, Z., etal. (2014).
Oral glycotoxins are a modifiable cause of dementia and the metabolic syn-
drome in mice and humans. Proc. Natl. Acad. Sci. U.S.A. 111, 4940–4945. doi:
10.1073/pnas.1316013111
Calió, M. L., Marinho, D. S., Ko, G. M., Ribeiro, R. R., Carbonel, A. F.,
Oyama, L. M., et al. (2014). Transplantation of bone marrow mesenchymal
stem cells decreases oxidative stress, apoptosis, and hippocampal damage in
brain of a spontaneous stroke model. Free Radic. Biol. Med. 70, 141–154. doi:
10.1016/j.freeradbiomed.2014.01.024
Cekanaviciute, E., Fathali, N., Doyle, K. P., Williams, A. M., Han, J., and Buckwal-
ter, M. S. (2014). Astrocytic transforming growth factor-beta signaling reduces
subacute neuroinflammation after stroke in mice. Glia 62, 1227–1240. doi:
10.1002/glia.22675
Ceruti, S., Villa, G., Genovese, T., Mazzon, E., Longhi, R., Rosa, P., et al. (2009).
The P2Y-like receptor GPR17 as a sensor of damage and a new potential tar-
get in spinal cord injury. Brain 132(Pt 8), 2206–2218. doi: 10.1093/brain/
awp147
Chen, C. J., Kono, H., Golenbock, D., Reed, G., Akira, S., and Rock, K. L. (2007).
Identification of a key pathway required for the sterile inflammatory response
triggered by dying cells. Nat. Med. 13, 851–856. doi: 10.1038/nm1603
Cho, S., Park, E. M., Febbraio, M., Anrather, J., Park, L., Racchumi, G., et al.
(2005). The class B scavenger receptor CD36 mediates free radical produc-
tion and tissue injury in cerebral ischemia. J. Neurosci. 25, 2504–2512. doi:
10.1523/JNEUROSCI.0035-05.2005
Chun, K. H., and Seong, S. Y. (2010). CD14 but not MD2 transmit signals from
DAMP. Int. Immunopharmaol. 10, 98–106. doi: 10.1016/j.intimp.2009.10.002
Clemens, J. A., Stephenson, D. T., Smalstig, E. B., Roberts, E. F., Johnstone, E.
M., Sharp, J. D., etal. (1996). Reactive glia express cytosolic phospholipase A2
after transient global forebrain ischemia in the rat. Stroke 27, 527–535. doi:
10.1161/01.STR.27.3.527
Cuartero, M. I., Ballesteros, I., Moraga, A., Nombela, F., Vivancos, J., Hamilton,
J. A., et al. (2013). N2 neutrophils, novel players in brain inflammation after
stroke: modulation by the PPARg agonist rosiglitazone. Stroke 44, 3498–3508.
doi: 10.1161/STROKEAHA.113.002470
Denes, A., Wilkinson, F., Bigger, B., Chu, M., Rothwell, N. J., and Allan, S. M.
(2013). Central and haematopoietic interleukin-1 both contribute to ischaemic
brain injury. Dis. Model. Mech. 6, 1043–1048. doi: 10.1242/dmm.011601
Desestret, V., Riou, A., Chauveau, F., Cho, T. H., Devillard, E., Marinescu, M.,
et al. (2013). In vitro and in vivo models of cerebral ischemia show discrep-
ancy in therapeutics effects of M2 macrophages. PLoS ONE. 8:e67063. doi:
10.1371/journal.pone.0067063
Dirnagl, U., Iadecola, C., and Moskowitz, M. A. (1999). Pathobiology of ischaemic
stroke: an integrated view. Trends Neurosci. 22, 391–397. doi: 10.1016/S0166-
2236(99)01401-0
Estévez, A. G., Sahawneh, M. A., Lange, P. S., Bae, N., Egea, M., and Ratan,
R. R. (2006). Arginase 1 regulation of nitric oxide production is key to sur-
vival of trophic factor-deprived motor neurons. J. Neurosci. 26, 8512–8516. doi:
10.1523/JNEUROSCI.0728-06.2006
Fann, D. Y., Lee, S. Y., Manzanero, S., Tang, S. C., Gelderblom, M., Chunduri,
P., et al. (2013). Intravenous immunoglobulin suppresses NLRP1 and NLRP3
inflammasome-mediated neuronal death in ischemic stroke. Cell Death Dis. 4,
e790. doi: 10.1038/cddis.2013.326
Franklin, B. S., Bossaller, L., De Nardo, D., Ratter, J. M., Stutz, A., Engels,
G., et al. (2014). The adaptor ASC has extracellular and ‘prionoid’ activi-
ties that propagate inflammation. Nat. Immunol. 15, 727–737. doi: 10.1038/
ni.2913
Furuyashiki, T., and Narumiya, S. (2011). Stress responses: the contribution of
prostaglandin E2and its receptors. Nat. Rev. Endocrinol. 7, 163–175. doi:
10.1038/nrendo.2010.194
Gao, S., Zhang, R., Greenberg, M. E., Sun, M., Chen, X., Levison, B. S.,
et al. (2006). Phospholipid hydroxyalkenals, a subset of recently discovered
endogenous CD36 ligands, spontaneously generate novel furan-containing phos-
pholipids lacking CD36 binding activity in vivo.J. Biol. Chem. 281, 31298–31308.
doi: 10.1074/jbc.M604039200
Gorina, R., Lyck, R., Vestweber, D.,and Engelhardt, B. (2014). β2 integrin-mediated
crawling on endothelial ICAM-1 and ICAM-2 is a prerequisite for transcellular
neutrophil diapedesis across the inflamed blood–brain barrier. J. Immunol. 192,
324–337. doi: 10.4049/jimmunol.1300858
Gudi, V., Škuljec, J., Yildiz, Ö., Frichert, K., Skripuletz, T., Moharregh-Khiabani,
D., etal. (2011). Spatial and temporal profiles of growth factor expression during
CNS demyelination reveal the dynamics of repair priming. PLoS ONE 6:e22623.
doi: 10.1371/journal.pone.0022623
Guo, L., Ge, J., Wang, S., Zhou, Y., Wang, X., and Wu, Y. (2013). A novel method for
efficient delivery of stem cells to the ischemic brain. Stem Cell Res. Ther. 4, 116.
doi: 10.1186/scrt327
Guo, Z., Zhang, L., Wu, Z., Chen, Y., Wang, F., and Chen, G. (2014). In Vivo
direct reprogramming of reactive glial cells into functional neurons after brain
injury and in an Alzheimer’s disease model. Cell Stem Cell 14, 188–202. doi:
10.1016/j.stem.2013.12.001
Haider, L., Fischer, M. T., Frishcer, J. M., Bauer, J., Höftberger, R., Botond, G.,
et al. (2011). Oxidative damage in multiple sclerosis lesions. Brain 134(Pt 7),
1914–1924. doi: 10.1093/brain/awr128
Hill, J. J., Jin, K., Mao, X. O., Xie, L., and Greenberg, D. A. (2012). Intracerebral
chondroitinase ABC and heparan sulfate proteoglycan glypican improveoutcome
from chronic stroke in rats. Proc. Natl. Acad. Sci. U.S.A. 109, 9155–9160. doi:
10.1073/pnas.1205697109
Ho, P. P., Kanter, J. L., Johnson, A. M., Srinagesh, H. K., Chang, E. J., Purdy,
T. M., et al. (2012). Identification of naturally occurring fatty acids of the
myelin sheath that resolve neuroinflammation. Sci. Transl. Med. 4, 137ra73. doi:
10.1126/scitranslmed.3003831
Hu, X., Li, P., Guo, Y., Wang, H., Leak, R. K., Chen, S., et al. (2012).
Microglia/macrophage polarization dynamics reveal novel mechanism of
injury expansion after focal cerebral ischemia. Stroke 43, 3063–3070. doi:
10.1161/STROKEAHA.112.659656
Hyakkoku, K., Hamanaka, J., Tsuruma, K., Shimazawa, M., Tanaka, H., Uematsu,
S., et al. (2010). Toll-like receptor 4 (TLR4), but not TLR3 or TLR9, knock-out
mice have neuroprotective effects against focal cerebral ischemia. Neuroscience
171, 258–267. doi: 10.1016/j.neuroscience.2010.08.054
Iadecola, C., andAnrather, J. (2011). The immunology of stroke: from mechanisms
to translation. Nat. Med. 17, 796–808. doi: 10.1038/nm.2399
Ikeda, N., Nonoguchi, N., Zhao, M. Z., Watanabe, T., Kajimoto, Y., Furutama,
D., et al. (2005). Bone marrow stromal cells that enhanced fibroblast growth
factor-2 secretion by herpes simplex virus vector improve neurological out-
come after transient focal cerebral ischemia in rats. Stroke 36, 2725–2730. doi:
10.1161/01.STR.0000190006.88896.d3
Ikeno, Y., Konno, N., Cheon, S. H., Bolchi, A., Ottonello, S., Kitamoto, K., etal.
(2005). Secretory phospholipases A2 induce neurite outgrowth in PC12 cells
through lysophosphatidylcholine generation and activation of G2A receptor. J.
Biol. Chem. 280, 28044–28052. doi: 10.1074/jbc.M503343200
Frontiers in Cellular Neuroscience www.frontiersin.org October 2014 |Volume 8 |Article 319 |6
Shichita et al. Inflammatory mechanisms after ischemic stroke
Iyer, S. S., He, Q., Janczy, J. R., Elliott, E. I., Zhong, Z., Olivier, A. K.,
et al. (2013). Mitochondrial cardiolipin is required for Nlrp3 inflammasome
activation. Immunity 39, 311–323. doi: 10.1016/j.immuni.2013.08.001
Kalladka, D., and Muir, K. W. (2014). Brain repair: cell therapy in stroke. Stem Cells
Cloning 7, 31–44.
Kawai, T., and Akira, S. (2010). The role of pattern-recognition receptors in
innate immunity: update on Toll-like receptors. Nat. Immunol. 11, 373–384.
doi: 10.1038/ni.1863
Konoeda, F., Shichita, T., Yoshida, H., Sugiyama, Y., Muto, G., Hasegawa, E., etal.
(2010). Therapeutic effect of IL-12/23 and their signaling pathway blockade on
brain ischemia model. Biochem. Biophys. Res. Commun. 402, 500–506. doi:
10.1016/j.bbrc.2010.10.058
Kooijman, R., Sarre, S., Michotte, Y., and De Keyser, J. (2009). Insulin-like growth
factor I: a potential neuroprotectivecompound for the treatment of acute ischemic
stroke? Stroke 40, e83–e88. doi: 10.1161/STROKEAHA.108.528356
Kuang, X., Wang, L. F., Yu, L., Li, Y. J., Wang, Y. N., He, Q.,
et al. (2014). Ligustilide ameliorates neuroinflammation and brain injury
in focal cerebral ischemia/reperfusion rats: involvement of inhibition of
TLR4/peroxiredoxin 6 signaling. Free Radic. Biol. Med. 71, 165–175. doi:
10.1016/j.freeradbiomed.2014.03.028
Lalancette-Hébert, M., Gowing, G., Simard, A., Weng, Y. C., and Kriz, J.
(2007). Selective ablation of proliferating microglial cells exacerbates ischemic
injury in the brain. J. Neurosci. 27, 2596–2605. doi: 10.1523/JNEUROSCI.5360
-06.2007
Lalancette-Hébert, M., Swarup, V., Beaulieu, J. M., Bohacek, I., Abdelhamid, E.,
Weng, Y. C., et al. (2012). Galectin-3 is required for resident microglia activation
and proliferation in response to ischemic injury. J. Neurosci. 32, 10383–10395.
doi: 10.1523/JNEUROSCI.1498-12.2012
Lecker, R. R., Soldner, F., Velasco, I., Gavin, D. K., Androutsellis-Theotokis, A., and
McKay, R. D. (2007). Long-lasting regeneration after ischemia in the cerebral
cortex. Stroke 38, 153–161. doi: 10.1161/01.STR.0000252156.65953.a9
Leung, P. Y., Stevens, S. L., Packard, A. E., Lessov, N. S., Yang, T., Conrad, V. K.,
et al. (2012). Toll-like receptor 7 preconditioning induces robust neuroprotection
against stroke by a novel type I interferon-mediated mechanism. Stroke 43, 1383–
1389. doi: 10.1161/STROKEAHA.111.641522
Liang, X., Lin, L., Woodling, N. S., Wang, Q., Anacker, C., Pan, T., et al. (2011).
Signaling via the prostaglandin E2 receptor EP4 exerts neuronal and vascular
protection in a mouse model of cerebral ischemia. J. Clin. Invest. 121, 4362–4371.
doi: 10.1172/JCI46279
Lo, E. H. (2010). Degeneration and repair in central nervous system disease. Nat.
Med. 16, 1205–1209. doi: 10.1038/nm.2226
Loser, K., Vogl, T., Voskort, M., Leuken, A., Kupas, V., Nacken, W., et al. (2010).
The Toll-like receptor 4 ligands Mrp8 and Mrp14 are crucial in the develop-
ment of autoreactive CD8+T cells. Nat. Med. 16, 713–717. doi: 10.1038/nm.
2150
Mabuchi, T., Kitagawa, K., Ohtsuki, T., Kuwabara, K., Yagita, Y., Yanagihara, T.,
et al. (2000). Contribution of microglia/macrophages to expansion of infarction
and response of oligodendrocytes after focal cerebral ischemia in rats. Stroke 31,
1735–1743. doi: 10.1161/01.STR.31.7.1735
Maeda, A., and Fadeel, B. (2014). Mitochondria released by cells undergoing
TNF-α-induced necroptosis act as danger signals. Cell Death Dis. 5, e1312. doi:
10.1038/cddis.2014.277
Marcheselli, V. L., Hong, S., Lukiw, W. J., Tian, X. H., Gronert, K., Musto, A.,
et al. (2003). Novel docosanoids inhibit brain ischemia-reperfusion-mediated
leukocyte infiltration and pro-inflammatory gene expression. J. Biol. Chem. 278,
43807–43817. doi: 10.1074/jbc.M305841200
Martinon, F., Burns, K., and Tschopp, J. (2002). The inflammasome: a molecular
platform triggering activation of inflammatory caspases and processing of proIL-
β.Mol. Cell 10, 417–426. doi: 10.1016/S1097-2765(02)00599-3
Matt, U., Sharif, O., Martins, R., Furtner, T., Langeberg, L., Gawish, R., et al. (2013).
WAVE1 mediates suppression of phagocytosis by phospholipid-derived DAMPs.
J. Clin. Invest. 123, 3014–3024. doi: 10.1172/JCI60681
McCullough, L., Wu, L., Haughey, N., Liang, X., Hand, T., Wang, Q., et al. (2004).
Neuroprotective function of the PGE2EP2 receptor in cerebral ischemia. J.
Neurosci. 24, 257–268. doi: 10.1523/JNEUROSCI.4485-03.2004
Meisel, C., and Meisel, A. (2011). Suppressing immunosuppression after
stroke. N. Engl. J. Med. 365, 2134–2136. doi: 10.1056/NEJMcibr11
12454
Miller, Y. I., Choi, S. H., Wiesner, P., Fang, L., Harkewicz, R., Hartvigsen, K.,
et al. (2011). Oxidation-specific epitopes are danger-associated molecular pat-
terns recognized by pattern recognition receptors of innate immunity. Circ. Res.
108, 235–248. doi: 10.1161/CIRCRESAHA.110.223875
Miyata, T., Oda, O., Inagi, R., Iida, Y., Araki, N., Yamada, N., et al. (1993). beta
2-Microglobulin modified with advanced glycation end products is a major com-
ponent of hemodialysis-associated amyloidosis. J. Clin. Invest. 92, 1243–1252.
doi: 10.1172/JCI116696
Moskowitz, M. A., Lo, E. H., and Iadecola, C. (2010). The science of
stroke: mechanisms in search of treatments. Neuron 67, 181–198. doi:
10.1016/j.neuron.2010.07.002
Muhammad, S., Barakat, W., Stoyanov, S., Murikinati, S., Yang, H., Tracey, K. J.,
et al. (2008). The HMGB1 receptor RAGE mediates ischemic brain damage. J.
Neurosci. 28, 12023–12031. doi: 10.1523/JNEUROSCI.2435-08.2008
Muralikrishna Adibhatla, R., and Hatcher, J. F. (2006). Phospholipase A2, reactive
oxygen species, and lipid peroxidation in cerebral ischemia. Free Radic. Biol. Med.
40, 376–387. doi: 10.1016/j.freeradbiomed.2005.08.044
Ness, J. K., Scaduto, R. C. Jr., and Wood, T. L. (2004). IGF-I prevents glutamate-
mediated bax translocation and cytochrome C release in O4+oligodendrocyte
progenitors. Glia 46, 183–194. doi: 10.1002/glia.10360
Ooboshi, H., Ibayashi, S., Shichita, T., Kumai, Y., Takada, J., Ago, T.,
et al. (2006). Postischemic gene transfer of interleukin-10 protects against
both focal and global brain ischemia. Circulation 111, 913–919. doi:
10.1161/01.CIR.0000155622.68580.DC
Pál, G., Vincze, C., Renner, É., Wappler, E. A., Nagy, Z., Lovas, G., et al. (2012).
Time course, distribution and cell types of induction of transforming growth
factor betas following middle cerebral artery occlusion in the rat brain. PLoS
ONE 7:e46731. doi: 10.1371/journal.pone.0046731
Qiang, X., Yang, W. L., Wu, R., Zhou, M., Jacob, A., Dong, W., et al. (2013).
Cold-inducible RNA-binding protein (CIRP) triggers inflammatory responses in
hemorrhagic shock and sepsis. Nat. Med. 19, 1489–1495. doi: 10.1038/nm.3368
Qiu, J., Nishimura, M., Wang, Y., Sims, J. R., Qiu, S., Savitz, S. I., et al. (2008). Early
release of HMGB-1 from neurons after the onset of brain ischemia. J. Cereb. Blood
Flow Metab. 28, 927–938. doi: 10.1038/sj.jcbfm.9600582
Qu, W. S., Wang, Y. H., Ma, J. F., Tian, D. S., Zhang, Q., Pan, D. J., et al.
(2011). Galectin-1 attenuates astrogliosis-associated injuries and improves recov-
ery of rats following focal cerebral ischemia. J. Neurochem. 116, 217–226. doi:
10.1111/j.1471-4159.2010.07095.x
Quintá, H. R., Pasquini, J. M., Rabinovich, G. A., and Pasquini, L. A. (2014). Glycan-
dependent binding of galectin-1 to neuropilin-1 promotes axonal regeneration
after spinal cord injury. Cell Death. Differ. 21, 941–955. doi:10.1038/cdd.2014.14
Reed-Geaghan, E. G., Savage, J. C., Hise, A. G., and Landreth, G. E. (2009). CD14
and toll-like receptors 2 and 4 are required for fibrillar Aβ-stimulated microglial
activation. J. Neurosci. 29, 11982–11992. doi: 10.1523/JNEUROSCI.3158-09.2009
Serhan, C. N. (2014). Pro-resolving lipid mediators are leads for resolution
physiology. Nature 510, 92–101. doi: 10.1038/nature13479
Shanta, S. R., Choi, C. S., Lee, J. H., Shin, C. Y., Kim, Y. J., Kim, K. H.,
et al. (2012). Global changes in phospholipids identified by MALDI MS in
rats with focal cerebral ischemia. J. Lipid Res. 53, 1823–1831. doi: 10.1194/jlr.
M022558
Shichita, T., Hasegawa, E., Kimura, A., Morita, R., Sakaguchi, R., Takada, I.,
et al. (2012). Peroxiredoxin family proteins are key initiators of post-ischemic
inflammation in the brain. Nat. Med. 18, 911–917. doi: 10.1038/nm.2749
Shichita, T.,Sugiyama, Y., Ooboshi, H., Sugimori, H., Nakagawa, R., Takada, I., et al.
(2009). Pivotal role of cerebral interleukin-17-producing γδT cells in the delayed
phase of ischemic brain injury. Nat. Med. 215, 946–950. doi: 10.1038/nm.1999
Smirkin, A., Matsumoto, H., Takahashi, H., Inoue, A., Tagawa, M., Ohue, S., et al.
(2010). Iba1+/NG2+macrophage-like cells expressing a variety of neuroprotec-
tive factors ameliorate ischemic damage of the brain. J. Cereb. Blood Flow Metab.
30, 603–615. doi: 10.1038/jcbfm.2009.233
Spohr,T. C., Dezonne, R. S., Rehen, S. K., and Gomes, F. C. (2011). Astrocytes treated
by lysophosphatidic acid induce axonal outgrowth of cortical progenitors through
extracellular matrix protein and epidermal growth factor signaling pathway. J.
Neurochem. 119, 113–123. doi: 10.1111/j.1471-4159.2011.07421.x
Starossom, S. C., Mascanfroni, I. D., Imitola, J., Cao, L., Raddassi, K., Hernan-
dez, S. F., et al. (2012). Galectin-1 deactivates classically activated microglia and
protects from inflammation-induced neurodegeneration. Immunity 37, 249–263.
doi: 10.1016/j.immuni.2012.05.023
Frontiers in Cellular Neuroscience www.frontiersin.org October 2014 |Volume 8 |Article 319 |7
Shichita et al. Inflammatory mechanisms after ischemic stroke
Stevens, S. L., Leung, P. Y., Vartanian, K. B., Gopalan, B., Yang, T., Simon,
R. P., et al. (2011). Multiple preconditioning paradigms converge on inter-
feron regulatory factor-dependent signaling to promote tolerance to ischemic
brain injury. J. Neurosci. 31, 8456–8463. doi: 10.1523/JNEUROSCI.0821-
11.2011
Stewart, C. R., Stuart, L. M., Wilkinson, K., van Gils, J. M., Deng, J., Halle,
A., et al. (2010). CD36 ligands promote sterile inflammation through assem-
bly of a Toll-like receptor 4 and 6 heterodimer. Nat. Immunol. 11, 155–161. doi:
10.1038/ni.1836
Sun, S., Sursal, T., Adibnia, Y., Zhao, C., Zheng, Y., Li, H., et al. (2013). Mitochon-
drial DAMPs increase endothelial permeability through neutrophil dependent
and independent pathways. PLoS ONE 8:e59989. doi: 10.1371/journal.pone.
0059989
Suzuki, Y., Hattori, K., Hamanaka, J., Murase, T., Egashira, Y., Mishiro,
K., et al. (2012). Pharmacological inhibition of TLR4-NOX4 signal protects
against neuronal death in transient focal ischemia. Sci. Rep. 2, 896. doi:
10.1038/srep00896
Suzuki, Y., Nakano, Y., Mishiro, K., Takagi, T., Tsuruma, K., Nakamura, M., etal.
(2013). Involvement of Mincle and Syk in the changes to innate immunity after
ischemic stroke. Sci. Rep. 3, 3177. doi: 10.1038/srep03177
Takeuchi, M., Yanase, Y., Matsuura, N., Yamagishi Si, S., Kameda,Y., Bucala, R., et al.
(2001). Immunological detection of a novel advanced glycation end-product.
Mol. Med. 7, 783–791.
Tang, S. C., Arumugam, T. V., Xu, X., Cheng, A., Mughal, M. R., Jo, D. G., etal.
(2007). Pivotal role for neuronal Toll-like receptors in ischemic brain injury
and functional deficits. Proc. Natl. Acad. Sci. U.S.A. 104, 13798–13803. doi:
10.1073/pnas.0702553104
Taniguchi, H., Anacker, C., Suarez-Mier, G. B., Wang, Q., and Andreasson, K.
(2011). Function of prostaglandin E2 EP receptors in the acute outcome of
rodent hypoxic ischemic encephalopathy. Neurosci. Lett. 504, 185–190. doi:
10.1016/j.neulet.2011.09.005
Tsai, S. Y., Segovia, J. A., Chang, T. H., Morris, I. R., Berton, M. T., Tessier, P. A.,
et al. (2014). DAMP molecule S100A9 acts as a molecular pattern to enhance
inflammation during influenza A virus infection: role of DDX21-TRIF-TLR4-
MyD88 pathway. PLoS Pathog. 10:e1003848. doi: 10.1371/journal.ppat.1003848
Uchida, K. (2013). Redox-derived damage-associated molecular patterns: lig-
and function of lipid peroxidation adducts. Redox Biol. 1, 94–96. doi:
10.1016/j.redox.2012.12.005
Walko, T. D. 3rd, Bola, R. A., Hong, J. D., Au, A. K., Bell, M. J., Kochanek, P.
M., et al. (2014). Cerebrospinal fluid mitochondrial DNA: a novel DAMP in
pediatric traumatic brain injury. Shock 41, 499–503. doi: 10.1097/SHK.00000000
00000160
Wei, Y., Yemisci, M., Kim,H. H., Yung, L. M., Shin, H. K., Hwang, S. K., et al. (2011).
Fingolimod provides long-term protection in rodent models of cerebral ischemia.
Ann. Neurol. 69, 119–129. doi: 10.1002/ana.22186
Weil, Z. M. (2012). Ischemia-induced hyperglycemia: consequences, neuroen-
docrine regulation, and a role for RAGE. Horm. Behav. 62, 280–285. doi:
10.1016/j.yhbeh.2012.04.001
Wenceslau, C. F.,McCarthy, C. G., Szasz, T., Spitler, K., Goulopoulou, S., Webb, R. C.,
et al. (2014). Mitochondrial damage-associated molecular patterns and vascular
function. Eur. Heart J. 35, 1172–1177. doi: 10.1093/eurheartj/ehu047
West, X. Z., Malinin, N. L., Merkulova, A. A., Tischenko, M., Kerr, B. A., Borden,
E. C., et al. (2010). Oxidative stress induces angiogenesis by activating TLR2 with
novel endogenous ligands. Nature 467, 972–976. doi: 10.1038/nature09421
Yanai, H., Ban, T., Wang, Z., Choi, M. K., Kawamura, T., Negishi, H.,
et al. (2009). HMGB proteins function as universal sentinels for nucleic-
acid-mediated innate immune responses. Nature 462, 99–103. doi: 10.1038/
nature08512
Yang, F., Wang, Z., Wei, X., Han, H., Meng, X., Zhang, Y., et al. (2014). NLRP3
deficiency ameliorates neurovascular damage in experimental ischemic stroke. J.
Cereb. Blood Flow Metab. 34, 660–667. doi: 10.1038/jcbfm.2013.242
Yang, Y., Liu, B., Dai, J., Srivastava, P. K., Zammit, D. J., Lefrançois, L., etal. (2007).
Heat shock protein gp96 is a master chaperone for Toll-like receptors and is
important in the innate function of macrophages. Immunity 26, 215–226. doi:
10.1016/j.immuni.2006.12.005
Yednock, T. A., Cannon, C., Fritz, L. C., Sanchez-Madrid, F., Steinman, L., and
Karin, N. (1992). Prevention of experimental autoimmune encephalomyeli-
tis by antibodies against α4β1 integrin. Nature 356, 63–66. doi: 10.1038/
356063a0
Zhang, J., Takahashi, H. K., Liu, K., Wake, H., Liu, R., Maruo, T., etal. (2011).
Anti-high mobility group box-1 monoclonal antibody protects the blood–brain
barrier from ischemia-induced disruption in rats. Stroke 42, 1420–1428. doi:
10.1161/STROKEAHA.110.598334
Zhang, Q., Raoof, M., Chen, Y., Sumi, Y., Sursal, T., Junger, W., et al. (2010). Cir-
culating mitochondrial DAMPs cause inflammatory responses to injury. Nature
464, 104–107. doi: 10.1038/nature08780
Zhong, Z., Zhai, Y., Liang, S., Mori, Y.,Han, R., Sutterwala, F. S., et al. (2013). TRPM2
links oxidative stress to NLRP3 inflammasome activation. Nat. Commun. 4, 1611.
doi: 10.1038/ncomms2608
Zhu, W., Fan, Y., Frenzel, T., Gasmi, M., Bartus, R. T., Young, W. L., etal.
(2008). Insulin growth factor-1 gene transfer enhances neurovascular remod-
eling and improves long-term stroke outcome in mice. Stroke 39, 1254–1261. doi:
10.1161/STROKEAHA.107.500801
Zhu, W., Fan, Y., Hao, Q., Shen, F., Hashimoto, T., Yang, G. Y., et al.
(2009). Postischemic IGF-1 gene transfer promotes neurovascular regenera-
tion after experimental stroke. J. Cereb. Blood Flow Metab. 29, 1528–1537. doi:
10.1038/jcbfm.2009.75
Ziegler, G., Freyer, D., Harhausen, D., Khojasteh, U., Nietfeld, W., and Trende-
lenburg, G. (2011). Blocking TLR2 in vivo protects against accumulation of
inflammatory cells and neuronal injury in experimental stroke. J. Cereb. Blood
Flow Metab. 31, 757–766. doi: 10.1038/jcbfm.2010.161
Conflict of Interest Statement: The authors declare that the research was conducted
in the absence of any commercial or financial relationships that could be construed
as a potential conflict of interest.
Received: 28 July 2014; accepted: 23 September 2014; published online: 14 October
2014.
Citation: Shichita T, Ito M and Yoshimura A (2014) Post-ischemic inflamma-
tion regulates neural damage and protection. Front. Cell. Neurosci. 8:319. doi:
10.3389/fncel.2014.00319
This article was submitted to the journal Frontiers in Cellular Neuroscience.
Copyright © 2014 Shichita, Ito and Yoshimura. This is an open-access article distributed
under the terms of the Creative Commons Attribution License (CC BY). The use, dis-
tribution or reproduction in other forums is permitted, provided the original author(s)
or licensor are credited and that the original publication in this journal is cited, in
accordance with accepted academic practice. No use, distribution or reproduction is
permitted which does not comply with these terms.
Frontiers in Cellular Neuroscience www.frontiersin.org October 2014 |Volume 8 |Article 319 |8
... 3.1.1 Pro-inflammatory cytokines and PSD (Fig. 1) In an ischemic stroke, pro-inflammatory cytokines, containing IL-1β, IL-6, and TNF-α, are released from glial cells and/or neurons and trigger the inflammation [45][46][47]. Cerebral ischemia-induced inflammation directly exacerbate the pathologies via injuring the blood-brain barrier (BBB) and promoting neuronal cell death [45,47]. Accumulating literatures have explored the relationship between the cytokines and PSD [48][49][50][51][52]. IL-18, a novel biomarker for PSD, is independently associated with depressive symptoms after stroke [48,52,53]. ...
... Pro-inflammatory cytokines and PSD (Fig. 1) In an ischemic stroke, pro-inflammatory cytokines, containing IL-1β, IL-6, and TNF-α, are released from glial cells and/or neurons and trigger the inflammation [45][46][47]. Cerebral ischemia-induced inflammation directly exacerbate the pathologies via injuring the blood-brain barrier (BBB) and promoting neuronal cell death [45,47]. Accumulating literatures have explored the relationship between the cytokines and PSD [48][49][50][51][52]. IL-18, a novel biomarker for PSD, is independently associated with depressive symptoms after stroke [48,52,53]. ...
... While increasing brain perfusion is not the only mechanism in which greater omentum plays an active role in stroke treatment, the immune ability and ability to produce neurotrophic factors may be the key factors as well. Studies have proved that peripheral macrophages have neuroprotective effects and can promote the repair of the central nervous system after the acute phase of stroke [67]. Owe to the immune capacity of the omentum, the displaced omentum may exhibit nerve restoration through this immunomodulatory pathway during stroke. ...
... HMGB1 is a key member of a family of non-histone DNA-binding proteins called high mobility group box (HMGB) family (Jiang et al., 2012, Tao et al., 2015. HMGB1 acts as a neuroinflammatory mediator, activating microglial cells in neurodegenerative diseases and the postischemic brain (Shichita et al., 2014). Under ischemic conditions, HMGB-1 is released from its chromatin-bound state from the nucleus to the extracellular space, where it links to its receptor sites and TLRs. ...
Article
Full-text available
Inflammatory responses and oxidative stress contribute to the pathogenesis of brain ischemia/reperfusion (IR) injury. Naturally occurring bioflavonoids possess antioxidant and anti-inflammatory properties. The phytochemicals of Juniperus sabina L., known as “Abhal” in Saudi Arabia, have been studied and cupressuflavone (CUP) has been isolated as the major bioflavonoid. This study aimed to investigate the neuroprotective potential of CUP in reducing brain IR damage in rats and to understand probable mechanisms. After 60 min of inducing cerebral ischemia by closing the left common carotid artery (CCA), blood flow was restored to allow reperfusion. The same surgical procedure was performed on sham-operated control rats, excluding cerebral IR. CUP or vehicle was given orally to rats for 3 days prior to ischemia induction and for a further 3 days following reperfusion. Based on the findings of this study, compared to the IR control group, CUP-administered group demonstrated reduced neurological deficits, improved motor coordination, balance, and locomotor activity. Additionally, brain homogenates of IR rats showed a decrease in malondialdehyde (MDA) level, an increase in reduced glutathione (GSH) content, and an increase in catalase (CAT) enzyme activity following CUP treatment. CUP suppressed neuro-inflammation via reducing serum inflammatory cytokine levels, particularly those of tumor necrosis factor alpha (TNF-α), interleukin-6 (IL-6), and interleukin-1 beta (IL-1β) and enhancing the inflammatory cytokine levels, such as Nuclear factor kappa- B (NF-κB), TANK-binding kinase-1 (TBK1), and interferon beta (IFN-β) in brain tissues. Furthermore, CUP ameliorated the histological alterations in the brain tissues of IR rats. CUP significantly suppressed caspase-3 expression and downregulated the Toll-like receptor 4 (TLR4)/NF-κB signaling pathway as a result of suppressing High mobility group box 1 (HMGB1). To our knowledge, this is the first study to document the neuroprotective properties of CUP. Thus, the study findings revealed that CUP ameliorates IR–induced cerebral injury possibly by enhancing brain antioxidant contents, reducing serum inflammatory cytokine levels, potentiating the brain contents of TBK1 and IFN-β and suppressing the HMGB1/TLR-4 signaling pathway. Hence, CUP may serve as a potential preventive and therapeutic alternative for cerebral stroke.
... 21 Moreover, activated microglial cells produce various mediators, including inducible nitric oxide synthase (iNOS), nitric oxide (NO), 27,28 pro-inflammatory cytokines (such as TNF-a), and anti-inflammatory cytokines (such as TGF-a), 29 collectively leading to complex immune reactions in the brains of patients with stroke, 16,[30][31][32] including reactions with contradictory effects. [33][34][35] Furthermore, there are numerous receptor-ligand pairs between microglial and immune cells, indicating the central role of microglial cells in immune responses, 21 as well as their critical involvement in neuroinflammatory-related diseases. 36 PRDX1 is a member of the peroxidase family that plays a crucial role in protecting cells from oxidative stress by detoxifying peroxides. ...
Article
Full-text available
Background Acute ischemic stroke (AIS) is a common cerebrovascular event associated with high incidence, disability, and poor prognosis. Studies have shown that various cell types, including microglia, astrocytes, oligodendrocytes, neurons, and neutrophils, play complex roles in the early stages of AIS and significantly affect its prognosis. Thus, a comprehensive understanding of the mechanisms of action of these cells will be beneficial for improving stroke prognosis. With the rapid development of single‐cell sequencing technology, researchers have explored the pathophysiological mechanisms underlying AIS at the single‐cell level. Method We systematically summarize the latest research on single‐cell sequencing in AIS. Result In this review, we summarize the phenotypes and functions of microglia, astrocytes, oligodendrocytes, neurons, neutrophils, monocytes, and lymphocytes, as well as their respective subtypes, at different time points following AIS. In particular, we focused on the crosstalk between microglia and astrocytes, oligodendrocytes, and neurons. Our findings reveal diverse and sometimes opposing roles within the same cell type, with the possibility of interconversion between different subclusters. Conclusion This review offers a pioneering exploration of the functions of various glial cells and cell subclusters after AIS, shedding light on their regulatory mechanisms that facilitate the transformation of detrimental cell subclusters towards those that are beneficial for improving the prognosis of AIS. This approach has the potential to advance the discovery of new specific targets and the development of drugs, thus representing a significant breakthrough in addressing the challenges in AIS treatment.
... Corr et al. have shown that one of these DAMPs is damaged myelin (Corr et al., 2016). These DAMPs skew microglia and macrophages toward pro-inflammatory phenotypes, as evidenced by their secretion of TNFα, IL-6, IL-1β, IFN-γ, and other mediators that increase cellular damage and activate pro-inflammatory immune cells (Shichita et al., 2014), (Zbesko et al., 2018). These myeloid cells also secrete reactive oxygen species (ROS) and matrix metalloproteinases (MMPs) (Jayaraj et al., 2023). ...
Article
Full-text available
Inflammation is a crucial part of the healing process after an ischemic stroke and is required to restore tissue homeostasis. However, the inflammatory response to stroke also worsens neurodegeneration and creates a tissue environment that is unfavorable to regeneration for several months, thereby postponing recovery. In animal models, inflammation can also contribute to the development of delayed cognitive deficits. Myeloid cells that take on a foamy appearance are one of the most prominent immune cell types within chronic stroke infarcts. Emerging evidence suggests that they form as a result of mechanisms of myelin lipid clearance becoming overwhelmed, and that they are a key driver of the chronic inflammatory response to stroke. Therefore, targeting lipid accumulation in foam cells may be a promising strategy for improving recovery. The aim of this review is to provide an overview of current knowledge regarding inflammation and foam cell formation in the brain in the weeks and months following ischemic stroke and identify targets that may be amenable to therapeutic intervention.
Article
Full-text available
Excessive inflammatory response following ischemic stroke (IS) injury is a key factor affecting the functional recovery of patients. The efferocytic clearance of apoptotic cells within ischemic brain tissue is a critical mechanism for mitigating inflammation, presenting a promising avenue for the treatment of ischemic stroke. However, the cellular and molecular mechanisms underlying efferocytosis in the brain after IS and its impact on brain injury and recovery are poorly understood. This study explored the roles of inflammation and efferocytosis in IS with bioinformatics. Three Gene Expression Omnibus Series (GSE) (GSE137482-3 m, GSE137482-18 m, and GSE30655) were obtained from NCBI (National Center for Biotechnology Information) and GEO (Gene Expression Omnibus). Differentially expressed genes (DEGs) were processed for GSEA (Gene Set Enrichment Analysis), GO (Gene Ontology Functional Enrichment Analysis), and KEGG (Kyoto Encyclopedia of Genes and Genomes) pathway analyses. Efferocytosis-related genes were identified from the existing literature, following which the relationship between Differentially Expressed Genes (DEGs) and efferocytosis-related genes was examined. The single-cell dataset GSE174574 was employed to investigate the distinct expression profiles of efferocytosis-related genes. The identified hub genes were verified using the dataset of human brain and peripheral blood sample datasets GSE56267 and GSE122709. The dataset GSE215212 was used to predict competing endogenous RNA (ceRNA) network, and GSE231431 was applied to verify the expression of differential miRNAs. At last, the middle cerebral artery (MCAO) model was established to validate the efferocytosis process and the expression of hub genes. DEGs in two datasets were significantly enriched in pathways involved in inflammatory response and immunoregulation. Based on the least absolute shrinkage and selection operator (LASSO) analyses, we identified hub efferocytosis-related genes (Abca1, C1qc, Ptx3, Irf5, and Pros1) and key transcription factors (Stat5). The scRNA-seq analysis showed that these hub genes were mainly expressed in microglia and macrophages which are the main cells with efferocytosis function in the brain. We then identified miR-125b-5p as a therapeutic target of IS based on the ceRNA network. Finally, we validated the phagocytosis and clearance of dead cells by efferocytosis and the expression of hub gene Abca1 in MCAO mice models.
Article
Full-text available
Ischemic cerebrovascular diseases are leading cause of mortality and disability worldwide. Given the need for a pharmacological treatment for these diseases, agmatine has gained great interest due to its neuroprotective properties. This article explores these properties of agmatine in ischemic events and their underlying mechanisms. Agmatine, considered as a neuromodulator, exerts its effects through its interaction with various molecular targets, including glutamate receptors, nitric oxide synthase, and metalloproteinases. Its ability to cross the blood-brain barrier and its role in neurotransmission processes postulate agmatine as a potential candidate for neuroprotection. Agmatine has a positive effect in the central nervous system to counteract excitotoxicity, oxidative stress, inflammation, alteration of the blood-brain barrier and energy disorders during ischemic events. This review describes the multiple interactions of agmatine within the ischemic cascade known to date, showing its ability to mitigate free radical formation, attenuate excitotoxicity, modulate inflammatory responses, stabilize the blood-brain barrier, and preserve mitochondrial function. These properties position agmatine as a promising therapeutic agent for ischemic cerebrovascular diseases.
Article
Microglia play key roles in the post‐ischemic inflammatory response and damaged tissue removal reacting rapidly to the disturbances caused by ischemia and working to restore the lost homeostasis. However, the modified environment, encompassing ionic imbalances, disruption of crucial neuron–microglia interactions, spreading depolarization, and generation of danger signals from necrotic neurons, induce morphological and phenotypic shifts in microglia. This leads them to adopt a proinflammatory profile and heighten their phagocytic activity. From day three post‐ischemia, macrophages infiltrate the necrotic core while microglia amass at the periphery. Further, inflammation prompts a metabolic shift favoring glycolysis, the pentose‐phosphate shunt, and lipid synthesis. These shifts, combined with phagocytic lipid intake, drive lipid droplet biogenesis, fuel anabolism, and enable microglia proliferation. Proliferating microglia release trophic factors contributing to protection and repair. However, some microglia accumulate lipids persistently and transform into dysfunctional and potentially harmful foam cells. Studies also showed microglia that either display impaired apoptotic cell clearance, or eliminate synapses, viable neurons, or endothelial cells. Yet, it will be essential to elucidate the viability of engulfed cells, the features of the local environment, the extent of tissue damage, and the temporal sequence. Ischemia provides a rich variety of region‐ and injury‐dependent stimuli for microglia, evolving with time and generating distinct microglia phenotypes including those exhibiting proinflammatory or dysfunctional traits and others showing pro‐repair features. Accurate profiling of microglia phenotypes, alongside with a more precise understanding of the associated post‐ischemic tissue conditions, is a necessary step to serve as the potential foundation for focused interventions in human stroke.
Article
Sonic Hedgehog (SHh) is a homology protein that is involved in the modeling and development of embryonic tissues. As SHh plays both protective and harmful roles in ischemia, any disruption in the transduction and regulation of the SHh signaling pathway causes ischemia to worsen. The SHh signal activation occurs when SHh binds to the receptor complex of Ptc-mediated Smoothened (Smo) (Ptc-smo), which initiates the downstream signaling cascade. This article will shed light on how pharmacological modifications to the SHh signaling pathway transduction mechanism alter ischemic conditions via canonical and non-canonical pathways by activating certain downstream signaling cascades with respect to protein kinase pathways, angiogenic cytokines, inflammatory mediators, oxidative parameters, and apoptotic pathways. The canonical pathway includes direct activation of interleukins (ILs), angiogenic cytokines like hepatocyte growth factor (HGF), platelet-derived growth factor (PDGF), vascular endothelial growth factor (VEGF), epidermal growth factor (EGF), and hypoxia-inducible factor alpha (HIF-), which modulate ischemia. The non-canonical pathway includes indirect activation of certain pathways like mTOR, PI3K/Akt, MAPK, RhoA/ROCK, Wnt/-catenin, NOTCH, Forkhead box protein (FOXF), Toll-like receptors (TLR), oxidative parameters such as GSH, SOD, and CAT, and some apoptotic parameters such as Bcl2. This review provides comprehensive insights that contribute to our knowledge of how SHh impacts the progression and outcomes of ischemic injuries.
Article
Cerebral ischemic stroke and glioma are the two leading causes of patient mortality globally. Despite physiological variations, 1 in 10 people who have an ischemic stroke go on to develop brain cancer, most notably gliomas. In addition, glioma treatments have also been shown to increase the risk of ischemic strokes. Stroke occurs more frequently in cancer patients than in the general population, according to traditional literature. Unbelievably, these events share multiple pathways, but the precise mechanism underlying their co-occurrence remains unknown. Transcription factors (TFs), the main components of gene expression programmes, finally determine the fate of cells and homeostasis. Both ischemic stroke and glioma exhibit aberrant expression of a large number of TFs, which are strongly linked to the pathophysiology and progression of both diseases. The precise genomic binding locations of TFs and how TF binding ultimately relates to transcriptional regulation remain elusive despite a strong interest in understanding how TFs regulate gene expression in both stroke and glioma. As a result, the importance of continuing efforts to understand TF-mediated gene regulation is highlighted in this review, along with some of the primary shared events in stroke and glioma.
Article
Full-text available
Microbes or danger signals trigger inflammasome sensors, which induce polymerization of the adaptor ASC and the assembly of ASC specks. ASC specks recruit and activate caspase-1, which induces maturation of the cytokine interleukin 1β (IL-1β) and pyroptotic cell death. Here we found that after pyroptosis, ASC specks accumulated in the extracellular space, where they promoted further maturation of IL-1β. In addition, phagocytosis of ASC specks by macrophages induced lysosomal damage and nucleation of soluble ASC, as well as activation of IL-1β in recipient cells. ASC specks appeared in bodily fluids from inflamed tissues, and autoantibodies to ASC specks developed in patients and mice with autoimmune pathologies. Together these findings reveal extracellular functions of ASC specks and a previously unknown form of cell-to-cell communication.
Article
Full-text available
Necrosis leads to the release of so-called damage-associated molecular patterns (DAMPs), which may provoke inflammatory responses. However, the release of organelles from dying cells, and the consequences thereof have not been documented before. We demonstrate here that mitochondria are released from cells undergoing tumor necrosis factor-α (TNF-α)-induced, receptor-interacting protein (RIP)1-dependent necroptosis, a form of programmed necrosis. The released, purified mitochondria were determined to be intact as they did not emit appreciable amounts of mitochondrial DNA (mtDNA). Pharmacological inhibition of dynamin-related protein 1 (Drp1) prevented mitochondrial fission in TNF-α-triggered cells, but this did not block necroptosis nor the concomitant release of mitochondria. Importantly, primary human macrophages and dendritic cells engulfed mitochondria from necroptotic cells leading to modulation of macrophage secretion of cytokines and induction of dendritic cell maturation. Our results show that intact mitochondria are released from necroptotic cells and suggest that these organelles act as bona fide danger signals.
Article
Full-text available
Assembly of the NLRP3 inflammasome activates caspase-1 and mediates the processing and release of the leaderless cytokine IL-1β and thereby serves a central role in the inflammatory response and in diverse human diseases. Here we found that upon activation of caspase-1, oligomeric NLRP3 inflammasome particles were released from macrophages. Recombinant oligomeric protein particles composed of the adaptor ASC or the p.D303N mutant form of NLRP3 associated with cryopyrin-associated periodic syndromes (CAPS) stimulated further activation of caspase-1 extracellularly, as well as intracellularly after phagocytosis by surrounding macrophages. We found oligomeric ASC particles in the serum of patients with active CAPS but not in that of patients with other inherited autoinflammatory diseases. Our findings support a model whereby the NLRP3 inflammasome, acting as an extracellular oligomeric complex, amplifies the inflammatory response.
Article
Full-text available
Microbes or danger signals trigger inflammasome sensors, which induce polymerization of the adaptor ASC and the assembly of ASC specks. ASC specks recruit and activate caspase-1, which induces maturation of the cytokine interleukin 1β (IL-1β) and pyroptotic cell death. Here we found that after pyroptosis, ASC specks accumulated in the extracellular space, where they promoted further maturation of IL-1β. In addition, phagocytosis of ASC specks by macrophages induced lysosomal damage and nucleation of soluble ASC, as well as activation of IL-1β in recipient cells. ASC specks appeared in bodily fluids from inflamed tissues, and autoantibodies to ASC specks developed in patients and mice with autoimmune pathologies. Together these findings reveal extracellular functions of ASC specks and a previously unknown form of cell-to-cell communication.
Article
Full-text available
Advances in our understanding of the mechanisms that bring about the resolution of acute inflammation have uncovered a new genus of pro-resolving lipid mediators that include the lipoxin, resolvin, protectin and maresin families, collectively called specialized pro-resolving mediators. Synthetic versions of these mediators have potent bioactions when administered in vivo. In animal experiments, the mediators evoke anti-inflammatory and novel pro-resolving mechanisms, and enhance microbial clearance. Although they have been identified in inflammation resolution, specialized pro-resolving mediators are conserved structures that also function in host defence, pain, organ protection and tissue remodelling. This Review covers the mechanisms of specialized pro-resolving mediators and omega-3 essential fatty acid pathways that could help us to understand their physiological functions.
Article
Full-text available
Astrocytes limit inflammation after CNS injury, at least partially by physically containing it within an astrocytic scar at the injury border. We report here that astrocytic transforming growth factor-beta (TGFβ) signaling is a second, distinct mechanism that astrocytes utilize to limit neuroinflammation. TGFβs are anti-inflammatory and neuroprotective cytokines that are upregulated subacutely after stroke, during a clinically accessible time window. We have previously demonstrated that TGFβs signal to astrocytes, neurons and microglia in the stroke border days after stroke. To investigate whether TGFβ affects astrocyte immunoregulatory functions, we engineered “Ast-Tbr2DN” mice where TGFβ signaling is inhibited specifically in astrocytes. Despite having a similar infarct size to wildtype controls, Ast-Tbr2DN mice exhibited significantly more neuroinflammation during the subacute period after distal middle cerebral occlusion (dMCAO) stroke. The peri-infarct cortex of Ast-Tbr2DN mice contained over 60% more activated CD11b+ monocytic cells and twice as much immunostaining for the activated microglia and macrophage marker CD68 than controls. Astrocytic scarring was not altered in Ast-Tbr2DN mice. However, Ast-Tbr2DN mice were unable to upregulate TGF-β1 and its activator thrombospondin-1 2 days after dMCAO. As a result, the normal upregulation of peri-infarct TGFβ signaling was blunted in Ast-Tbr2DN mice. In this setting of lower TGFβ signaling and excessive neuroinflammation, we observed worse motor outcomes and late infarct expansion after photothrombotic motor cortex stroke. Taken together, these data demonstrate that TGFβ signaling is a molecular mechanism by which astrocytes limit neuroinflammation, activate TGFβ in the peri-infarct cortex and preserve brain function during the subacute period after stroke. GLIA 2014
Conference Paper
Introduction: Recently, the role of danger-associated molecular pattern molecules (DAMPs, also known as alarmins) has been elucidated in the acute inflammatory response after traumatic brain injury (TBI). Previously, we reported that elevated cerebrospinal fluid (CSF) high mobility group box-1 (HMGB1) levels are associated with poor neurological outcome. Mitochondrial DNA (mtDNA), a novel DAMP, is released into the extracellular milieu after cell death and injury, and acts via Toll-like receptor 9 (TLR9), a pattern recognition receptor of the immune system that detects bacterial and viral DNA. Methods: All children with severe TBI who received CSF drainage for treatment of intracranial hypertension were included in this IRB approved study. CSF was continuously drained by gravity into a sterile buretrol and collected using sterile techniques. Using quantitative PCR, we measured free mtDNA in CSF samples from 36 children with TBI and 13 control subjects who had negative CSF from a lumbar puncture to rule out meningitis. We designed PCR primers that amplify bases between 421 and 781 encoding the mitochondrial gene cytochrome c oxidase I (RefSeq: NC_012920). The linearity of the quantitative assay was assessed by using a cloned plasmid DNA, which was serially diluted to prepare a series of calibrators with known concentrations (Chiu et al., 2003). All statistical analyses were performed using Mann-Whitney and Wilcoxon Rank-Sum test. Results: The mean age of the cohort was 6.3 yrs and 22 (61%) children were male. An initial Glasgow Coma Scale (GCS) score of <8 was documented in 70% of children with TBI; monitors were placed in the remaining 30% due to subsequent deterioration as GCS decreased to <= 8. CSF levels of mtDNA were significantly increased (log 2.38-fold change) in patients with TBI compared to the control population. In this preliminary data set, a positive correlation was noted between CSF levels of mtDNA, and another DAMP, HMGB1 at 24 h after TBI (rho = 0.5674, p<0.05), but there was no significant associations between mtDNA levels and age, sex and GCS or Glasgow Outcome Score. Conclusions: The elevated CSF concentrations of mtDNA, a novel DAMP, support the notion of mtDNA as an endogenous danger signal after TBI in children, and possible footprint of brain tissue necrosis given its association with CSF HMGB1. Our ongoing work is focused on identifying the pathways in the secondary injury response that are activated by this novel danger signal.
Article
Immune system activation occurs not only due to foreign stimuli, but also due to endogenous molecules. As such, endogenous molecules that are released into the circulation due to cell death and/or injury alarm the immune system that something has disturbed homeostasis and a response is needed. Collectively, these molecules are known as damage-associated molecular patterns (DAMPs). Mitochondrial DAMPs (mtDAMPs) are potent immunological activators due to the bacterial ancestry of mitochondria. Mitochondrial DAMPs are recognized by specific pattern recognition receptors of the innate immune system, some of which are expressed in the cardiovascular system. Cell death leads to release of mtDAMPs that may induce vascular changes by mechanisms that are currently not well understood. This review will focus on recently published evidence linking mtDAMPs and immune system activation to vascular dysfunction and cardiovascular disease. © 2014 Published on behalf of the European Society of Cardiology. All rights reserved. © The Author 2014. For permissions please email: [email protected] /* */