ArticlePDF Available

Reduced susceptibility to ischemic brain injury and N-methyl-D-aspartate-mediated neurotoxicity in cyclooxygenase-2-deficient mice

Authors:

Abstract and Figures

Cyclooxygenase-2 (COX-2), a prostanoid-synthesizing enzyme that contributes to the toxicity associated with inflammation, has recently emerged as a promising therapeutic target for several illnesses, ranging from osteoarthritis to Alzheimer's disease. Although COX-2 has also been linked to ischemic stroke, its role in the mechanisms of ischemic brain injury remains controversial. We demonstrate that COX-2-deficient mice have a significant reduction in the brain injury produced by occlusion of the middle cerebral artery. The protection can be attributed to attenuation of glutamate neurotoxicity, a critical factor in the initiation of ischemic brain injury, and to abrogation of the deleterious effects of postischemic inflammation, a process contributing to the secondary progression of the damage. Thus, COX-2 is involved in pathogenic events occurring in both the early and late stages of cerebral ischemia and may be a valuable therapeutic target for treatment of human stroke.
Content may be subject to copyright.
Reduced susceptibility to ischemic brain injury and
N
-methyl-D-aspartate-mediated neurotoxicity in
cyclooxygenase-2-deficient mice
Costantino Iadecola*
, Kiyoshi Niwa*, Shigeru Nogawa*, Xueren Zhao*, Masao Nagayama*, Eiichi Araki*,
Scott Morham
, and M. Elizabeth Ross*
*Center for Clinical and Molecular Neurobiology, Department of Neurology, University of Minnesota Medical School, Minneapolis, MN 55455;
and Myriad Genetics, Salt Lake City, UT 84108
Communicated by Philip Needleman, Monsanto Company, St. Louis, MO, November 21, 2000 (received for review July 27, 2000)
Cyclooxygenase-2 (COX-2), a prostanoid-synthesizing enzyme that
contributes to the toxicity associated with inflammation, has
recently emerged as a promising therapeutic target for several
illnesses, ranging from osteoarthritis to Alzheimer’s disease. Al-
though COX-2 has also been linked to ischemic stroke, its role in the
mechanisms of ischemic brain injury remains controversial. We
demonstrate that COX-2-deficient mice have a significant reduc-
tion in the brain injury produced by occlusion of the middle
cerebral artery. The protection can be attributed to attenuation
of glutamate neurotoxicity, a critical factor in the initiation of
ischemic brain injury, and to abrogation of the deleterious effects
of postischemic inflammation, a process contributing to the sec-
ondary progression of the damage. Thus, COX-2 is involved in
pathogenic events occurring in both the early and late stages of
cerebral ischemia and may be a valuable therapeutic target for
treatment of human stroke.
middle cerebral artery occlusion prostanoids cerebral blood
flow NS398 stroke
Stroke remains a major cause of death and disability world-
wide (1– 4). After many years of setbacks, effective treat-
ments for stroke, based on thrombolysis and restoration of flow,
have been developed (5, 6). However, these therapies can be
safely administered only in those patients who reach medical
attention during the first few hours after the onset of the stroke
(7, 8). Therefore, it would be important to develop treatments
that target downstream events in the ischemic cascade. Further-
more, it would be desirable to combine thrombolytic therapy
with other therapeutic strategies aimed at protecting the brain
from the residual ischemia (9). Immediately after induction of
ischemia, activation of glutamate receptors initiates the ischemic
cascade and contributes to the damage that occurs in the early
stages of cerebral ischemia (refs. 10 and 11; see ref. 12 for a
review). At later times after ischemia, the tissue damage con-
tinues to evolve (13), and inflammation and programmed cell
death are thought to be major factors in the progression of the
injury (refs. 14 –17; see ref. 18 for a review).
Cyclooxygenase (COX)-1 and COX-2 are enzymes involved in
the first step of the synthesis of prostanoids (see ref. 19 for a
review). COX-1 is expressed constitutively in many organs and
contributes to the synthesis of prostanoids involved in normal
cellular functions. COX-2 is thought to be an inducible enzyme
whose expression is up-regulated in pathological states, most
notably those associated with inf lammation (see ref. 20 for a
review). There is substantial evidence that COX-2 reaction
products are responsible for cytotoxicity in models of inflam-
mation (see ref. 21 for a review). Thus, COX-2 inhibitors are
potent antiinflammatory agents and have recently been intro-
duced in clinical practice with notable success (22, 23).
In brain, COX-2 is present in selected neurons (24–26), and
its expression is up-regulated in several neurological diseases,
including stroke, Alzheimer’s dementia, and seizures (24, 27–
29). In rodents as in humans, cerebral ischemia up-regulates
COX-2 in neurons, blood vessels, and inflammator y cells infil-
trating the injured brain (27, 28, 30). However, experimental
evidence linking COX-2 to the mechanisms of brain injury
associated with these conditions is lacking. For example, al-
though it is well established that cerebral ischemia increases
COX-2 expression in the damaged brain, studies using COX
inhibitors in models of focal cerebral ischemia have yielded
conflicting results (28, 31, 32). Furthermore, studies using
COX-2 inhibitors cannot exclude effects unrelated to COX-2,
such as modification of gene expression or activation of perox-
isome proliferator-activated receptors (33, 34). In view of the
increasing use of COX-2 inhibitors as antiinflammatory agents
in the elderly—a population with an increased risk for stroke—it
would be of great interest to define the role of COX-2 in ischemic
brain injury.
Mice with a null mutation of the COX-2 gene have been a
useful model for investigating the role of COX-2 in systemic
inflammation, thermoregulation, and cerebrovascular regula-
tion (refs. 35 and 36; see ref. 37 for a review). In the present
study, therefore, we used COX-2-deficient mice to gain further
insight into the role of COX-2 in ischemic brain injury.
Methods
Animals. COX-2-null mice were obtained from a colony estab-
lished at the University of Minnesota (35, 36). Mice (SV129
C57BL
6J) were back-crossed to C57BL
6J mice five or six
times and were studied at age 2–3 months. Experiments were
performed in age-matched littermates [wild-type (
), het-
erozygous (
), and homozygous (
)] to minimize con-
founding effects deriving from the genetic background of the
mice. The genotype of all COX-2 mice was determined by PCR
with the use of primers and methods described previously (35).
No alterations in the anatomy of large cerebral vessels were
noticed in COX-2-null mice. C57BL
6J mice were obtained
from The Jackson Laboratory.
Induction of Focal Cerebral Ischemia. Focal cerebral ischemia was
produced by occlusion of the middle cerebral artery (MCA) (38).
Mice were anesthetized with 2% halothane
100% oxygen. Body
temperature was maintained at 37 0.5°C by a thermostatically
Abbreviations: NMDA, N-methyl-D-aspartate; COX, cyclooxygenase; MCA, middle cerebral
artery; RT-PCR, reverse transcription–PCR; PBD, porphobilinogen deaminase; CBF, cerebral
blood flow; PGE2, prostaglandin E2.
To whom reprint requests should be addressed at: Department of Neurology, University
of Minnesota Medical School, Box 295, UMHC, 516 Delaware Street SE, Minneapolis, MN
55455. E-mail: iadec001@tc.umn.edu.
The publication costs of this article were defrayed in part by page charge payment. This
article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C.
§1734 solely to indicate this fact.
1294–1299
PNAS
January 30, 2001
vol. 98
no. 3
controlled infrared lamp. A 2-mm hole was drilled in the inferior
portion of the temporal bone to expose the left MCA. The MCA
was elevated and cauterized distal to the origin of the lenticu-
lostriate branches. Mice in which the MCA was exposed but not
occluded served as sham-operated controls. Wounds were su-
tured and mice were allowed to recover and were returned to
their cages. Rectal temperature was controlled until mice re-
gained full consciousness. Thereafter, rectal temperature was
measured daily until the time of sacrifice. There were no major
differences in rectal temperature among COX-2
,
,
and
mice before MCA occlusion or after ischemia. For
example, 48 h after MCA occlusion rectal temperature was
36.2 0.1°C in COX-2
, 36.6 0.1 in
, and 36.0 0.2
in
mice (P0.05, analysis of variance).
Reverse Transcription–PCR (RT-PCR). mRNA for COX-2 was de-
tected by RT-PCR as previously described (28). Mice were killed
24 h after ischemia (n4 per time point), and their brains were
removed. This time point was chosen because COX-2 mRNA
expression reaches a plateau 24 h after MCA occlusion in mice
(39). A 4-mm-thick coronal brain slice was cut at the level of the
optic chiasm, and samples including the infarcted cortex and the
corresponding area of the contralateral cortex were collected
and frozen in liquid nitrogen. Total RNA was extracted from the
tissue according to the method of Chomczynski and Sacchi (40).
RNA integrity was determined on denaturing formaldehyde
gels. Aliquots of total RNA (0.25
g) were used in the RT
reaction mixed with 0.5
g of oligo(dT) primer as directed
(18-mer; New England Biolabs). First-strand cDNA synthesis
was then carried out with the use of Moloney murine leukemia
virus reverse transcriptase (New England Biolabs) according to
the manufacturer’s instructions. After heating at 95°C for 10 min,
5
l from each RT reaction mixture was used for PCR ampli-
fication. Primers (0.2
M each) for the sequence of interest and
for porphobilinogen deaminase (PBD), a ubiquitously expressed
sequence, were used in a final volume of 50
l. COX-2 primers
were as follows: forward, 5-CCAGATGCTATCTTTGGG-
GAGAC-3; reverse, 5-GCTTGCATTGATGGTGGCTG-3,
which result in a PCR product of 249 bp. COX-1 primers were
as follows: forward, 5-GAATACCGAAAGAGGT TTG-
GCTTG-3; reverse, 5-TCATCTCCAGGGTAATCTGGCAC-
3, which yields a 374-bp product. PBD primers were as follows:
forward, 5-GCCACCACAGTCTCGGTCTGTATGCGAGC-
3; reverse, 5-TGTCCGGTAACGGCGGCGCGGCCA-
CAAC-3. The ‘‘hot start’’ method was used with the following
cycle parameters: 94°C, 15 s; 68°C, 30 s; 73°C, 20 s, for five cycles,
then 94°C, 15 s; 64°C, 30 s; 73°C, 20 s, for 35 cycles; and 73°C,
15 min. Reaction products were then separated on an 8%
polyacrylamide gel, stained with ethidium bromide, and photo-
graphed. Each set of PCRs included control samples run without
RNA or in which the RT step was omitted to ensure that PCR
products resulted from amplification from the COX-2 mRNA
rather than genomic DNA. The OD of the bands was determined
by a gel image analysis system (Molecular Analyst; Bio-Rad). In
some studies, measurements were normalized to the OD of the
PBD band used as an internal standard (28, 41).
Competitive RT-PCR was used to determine more accurately
the magnitude of mRNA induction (28). A deletion construct
was synthesized that consisted of the same sequence amplified
from the endogenous COX-2 message but missing an internal
79-nucleotide fragment. To generate the construct, a pair of
COX-2 primers was prepared: forward: 5-CCAGATGC-
TATCTTTGGGGAGAC-3; reverse: 5-GCTTGCATTGAT-
GGTGGCTG-3, which result in a 249-bp PCR product. The
PCR product was then digested with a restriction enzyme,
Sau3AI (New England Biolabs). The sample was then religated
at 14°C overnight with T4 DNA ligase (New England Biolabs),
and the ligase was inactivated at 65°C for 15 min. A 1-
l aliquot
of the religated sample was then amplified with the use of the
COX-2 forward and reverse primers described above. Products
were separated on a polyacrylamide gel, and a main product of
170 bp was excised from the gel. The construct was eluted from
the crushed gel with T.E. buffer (10 mM Tris
1 mM EDTA, pH
8.0) at 55°C for 4 h and purified for use in the competition assay.
The RT reaction mixtures (5
l each) of the animals killed 24 h
after ischemia (n4 per group) were coamplified with known
amounts of deletion construct (0.25–25 fg). The PCR products
were then separated on a gel, and the gel was stained with
ethidium bromide and photographed. The OD of the bands was
determined by image analysis. For data analysis, the log of the
OD ratio (COX-2
construct) was plotted as a function of the log
of the concentration of the construct and fitted by linear
regression analysis (28, 42). The 0 value of the log of the ratio
(COX-2
construct) (yaxis) represents the point at which the
COX-2 PCR product and the construct are present in equal
amounts. Therefore, the amount of the construct corresponding
to the 0 ratio (xaxis) represents the amount of the COX-2 PCR
product before PCR amplification (42, 43).
Prostaglandin E
2
(PGE
2
) Enzyme Immunoassay. Tissue concentration
of PGE
2
, a COX reaction product, was measured 24 h after MCA
occlusion or3hafterN-methyl-D-aspartate (NMDA) injection.
Fig. 1. COX-2 mRNA expression in the brain of COX-2-null mice 24 h after
MCA occlusion. (A) Representative gel illustrating COX-2 mRNA expression in
the ischemic cortex (s) and contralateral cortex (n) assessed by RT-PCR. The
ubiquitous sequence PBD was also studied and used as an internal control. std,
standards; bl, sample without the reverse transcriptase step. (B) COX-2 mRNA
expression assessed by competitive PCR. COX-2 expression is reduced in COX-2
mice and is absent in COX-2
mice (n4 per group; *,P0.05 from
contralateral; #, P0.05 from
stroke).
Iadecola et al. PNAS
January 30, 2001
vol. 98
no. 3
1295
PHARMACOLOGY
PGE
2
concentration was determined in the cerebral cortex
ipsilateral and contralateral to the lesion with the use of an
enzyme immunoassay kit (Cayman Chemicals, Ann Arbor, MI)
as previously described (28). Prostanoids were extracted with
100% methanol according to the method of Powell (see ref. 28),
and the PGE
2
concentration was determined spectrophotometri-
cally according to the instructions provided with the kit.
NMDA Microinjection in Neocortex. In halothane-anesthetized mice
the dura overlying the parietal cortex was exposed, and NMDA
(20 nmol in 200–300 nl of sterile 0.1 M PBS, pH 7.4) was injected
with a glass micropipette (tip 40 –50
m) connected to a micro-
injection device. The micropipette was inserted into the parietal
cortex at a site 1.5 mm caudal to bregma, 4.0 mm from the
midline, and 0.8 mm below the dural surface. The micropipette
was left in place for 10 min, to minimize back-f lux of NMDA, and
then removed. Mice were returned to their cages and allowed to
survive for 3 h for PGE
2
measurement and for 24 h for
determination of lesion volume.
Determination of Lesion Volume. Mice were killed 1 and 4 days
after MCA occlusion or 1 day after NMDA injection. Brains
were removed and frozen in cooled isopentane (30°C). Coro-
nal forebrain sections (thickness 30
m) were serially cut in
a cryostat, collected at 90-
m intervals, and stained with thionin
for determination of lesion volume by an image analyzer (MCID;
Imaging Research, St. Catherine’s, ON, Canada; ref. 38). In
studies of cerebral ischemia, infarct volume in cerebral cortex
was corrected for swelling to factor out the contribution of
ischemic edema to the total volume of the lesion (38, 44).
Monitoring of Cerebral Blood Flow (CBF). Techniques used for
monitoring CBF after MCA occlusion have been published (38).
Mice were anesthetized with halothane (maintenance 1%), and
the femoral artery and trachea were cannulated. Mice were
artificially ventilated with an oxygen–nitrogen mixture by a
mechanical ventilator (SAR-830; CWE, Ardmore, PA). The
oxygen concentration in the mixture was adjusted to maintain
arterial partial O
2
pressure between 120 and 140 mmHg. End-
tidal CO
2
was continuously monitored with a CO
2
analyzer
(Capstar-100; CWE) (38). CBF was monitored with two laser-
Doppler flow probes (Vasamedic, Minneapolis, MN) placed
through burr holes drilled in the center (3.5 mm lateral to the
midline and 1 mm caudal to bregma) and the periphery (1.5 mm
lateral to the midline and 1.7 mm rostral to lambda) of the
ischemic territory (38). The location of the probe was selected in
preliminary experiments to correspond to the region of brain
that is spared from infarction in COX-2-deficient mice. After
placement of the probes, the MCA was occluded and CBF was
monitored for 90 min. CBF data are expressed as a percentage
of the preocclusion value. Arterial pressure and blood gases did
not differ between COX-2
and
mice (10 min after
MCA occlusion: COX-2
: arterial pressure: 83 3 mmHg;
paCO
2
: 34.6 0.6 mmHg; arterial partial O
2
pressure: 125 6
mmHg; pH: 7.35 0.02; COX-2
: arterial pressure: 84
2; paCO
2
: 33.1 0.8; arterial partial O
2
pressure: 126 5; pH:
7.33 0.02).
Renal Function. Because COX-2-null mice develop renal disease
(35), plasma creatinine was measured by standard colorimetric
methods. Furthermore, histological analysis of the kidneys was
performed in paraffin-embedded sections stained with hema-
toxylin and eosin. Creatinine did not differ between COX-2
(0.18 0.07 mg
dl; n6) and
mice (0.18 0.05;
n6) (P0.05 from COX-2
). In COX-2
mice (n
6), creatinine was elevated (0.43 0.08; P0.05 from COX-2
) but still in the normal range (45). COX-2
mice had
hypertrophy of the juxtaglomerular apparatus and macula densa
but no alterations in the glomeruli or tubules.
Data Analysis. Data in the text and figures are expressed as
means SEM. Multiple comparisons were evaluated statisti-
cally by the analysis of variance and Tukey’s test. Two-group
comparisons were analyzed by the two-tailed Student’s ttest for
independent samples. For all procedures, probability values of
less than 0.05 were considered statistically significant.
Results
Postischemic COX-2 Expression Is Attenuated in COX-2-Null Mice.
First, we sought to establish whether the increase in COX-2
expression that occurs after cerebral ischemia is reduced in
COX-2-null mice. In COX-2
mice, COX-2 mRNA was
increased in the ischemic cortex 24 h after MCA occlusion by 2-
to 3-fold (Fig. 1A). The up-regulation was reduced in COX-2
(P0.05, analysis of variance) (Fig. 1B). COX-2 mRNA
was not detected in either the ischemic or the nonischemic cortex
in COX-2
mice. In contrast, expression of COX-1 mRNA
was comparable in COX-2
,
, and
mice (Fig.
2A). To determine whether the reduction in COX-2 mRNA
expression is associated with a reduction in COX-2 enzymatic
Fig. 2. COX-1 mRNA expression and PGE2concentration in the brain of
COX-2-null mice 24 h after MCA occlusion. (A) Representative gel illustrating
COX-1 mRNA expression assessed by RT-PCR 24 h after MCA occlusion. Similar
results were obtained in four separate groups of COX-2
,
, and
mice. COX-1
PBD OD ratios in the ischemic cortex of COX-2
,
, and
mice were 1.7 0.4, 1.6 0.2, and 1.8 0.5, respectively (n4 per
group; P0.05). Abbreviations are as in Fig. 1. (B) Effect of MCA occlusion on
PGE2concentration in the ischemic cortex (Stroke) and contralateral cortex.
The PGE2elevation in the ischemic cortex in COX-2
(n8) was attenu-
ated in COX-2
(n7) and abolished in COX-2
mice (n7) (*,P
0.05 from contralateral; #, P0.05 from
stroke).
1296
www.pnas.org Iadecola et al.
activity, the concentration of the COX reaction product PGE
2
was measured in the ischemic and nonischemic cortex 24 h after
MCA occlusion. PGE
2
concentration was significantly elevated
in the ischemic cortex of COX-2
mice (Fig. 2B). The
elevation was attenuated in COX-2
mice and absent in
COX-2
mice (Fig. 2B).
Ischemic Brain Injury Is Reduced in COX-2-Null Mice. We then studied
the brain injury produced by MCA occlusion in COX-2-null
mice. Mice were killed 96 h after ischemia, and the infarct
volume (mm
3
) was measured in brain sections stained with
thionin (38). The 96-h time point was selected on the basis of the
fact that, at this time, the damage resulting from postischemic
inflammation is fully expressed (38). Infarct volume was signif-
icantly smaller in COX-2-null mice than in wild-type littermates
(Fig. 3A). The reduction was greater in COX-2
(38
4%) than in
mice (20 3%) (Fig. 3A). These data
suggest that COX-2 contributes to ischemic brain injury.
To determine whether COX-2 is also involved in pathogenic
processes that occur in the initial stages of cerebral ischemia, we
studied infarct volume in mice killed 24 h after ischemia. At this
time the damage resulting from postischemic inflammation does
not contribute to tissue injury (38). For example, in mice lacking
inducible nitric oxide synthase, an enzyme that plays a critical
role in inflammatory responses, infarct volume is not reduced
24 h after ischemia, but only at 96 h (38). We found that in
COX-2-null mice infarct volume is also reduced 24 h after
ischemia (34 3%; Fig. 3B). These data suggest that COX-2
is also involved in pathogenic events occurring in the early stages
of cerebral ischemia.
Effect of MCA Occlusion on CBF in COX-2-Null Mice. The intensity of
the ischemic insult has a profound impact on the ensuing brain
damage (see ref. 46 for a review). We therefore studied the effect
of MCA occlusion on neocortical CBF in COX-2
and
mice. CBF was monitored both in the center of the ischemic
territory and in the peripheral region that is spared from
infarction in COX-2
mice. As illustrated in Fig. 3 Cand D,
the reduction in CBF in the center or periphery of the ischemic
territory did not differ between COX-2
and
mice
(P0.05). These data demonstrate that the reduction in CBF
in areas that are spared from infarction in COX-2
mice is
comparable to that observed in COX-2
mice. Therefore,
differences in the intensity of the ischemic insult cannot account
for the neuroprotection observed in COX-2-null mice.
NMDA-Mediated Injury
in Vivo
Is Attenuated in COX-2-Null Mice.
Glutamate receptors play a critical role in the initiation of
ischemic brain injury (12). Therefore, we sought to determine
whether the reduction in ischemic damage in COX-2-null mice
could be attributed to reduced susceptibility to glutamate re-
ceptor-mediated damage. The glutamate receptor agonist
NMDA was microinjected directly into the cerebral cortex, and
injury volume was determined in thionin-stained brain sections
24 h later. At this time after NMDA microinjection, there is no
histological evidence of inf lammation in the area of the lesion
(data not shown). In normal mice, NMDA microinjection in-
creased the local concentration of PGE
2
and produced a well-
defined neocortical lesion (Fig. 4 Aand B). The PGE
2
elevation
was attenuated by treatment with the COX-2 inhibitor NS398 (20
mg
kg i.p.; 1 h before NMDA), demonstrating that COX-2 was
responsible for the increase in this prostaglandin (Fig. 4A).
NS398 administration (20 mg
kg i.p.; 1 h before and1hafter
NMDA) also reduced the volume of the lesion (Fig. 4 Cand D).
In addition, the NMDA-induced damage was markedly attenu-
ated in COX-2
mice (Fig. 5). Administration of NS398
reduced injury volume in COX-2
but not in COX-2
mice (Fig. 5C), attesting to the specificity of the effect of NS398
on COX-2. Thus the brain damage produced by the activation of
glutamate receptors is attenuated by COX-2 inhibition and in
COX-2-null mice.
Fig. 3. (A) Infarct size in COX-2
(n6),
(n7), and
(n8) null mice 96 h after MCA occlusion (MCAO). Ctx (E.C.), cerebral cortical infarct
corrected for swelling. (*,P0.05 from COX-2
and
mice; #, P0.05 from COX-2
and
mice). (B) Infarct size in COX-2-null mice 24 h after
MCAO (n6 per group; *,P0.05 from COX-2
mice). (Cand D) CBF reduction in the center (C) and periphery (D) of the ischemic region in COX-2
and
mice after MCAO (n6 per group).
Iadecola et al. PNAS
January 30, 2001
vol. 98
no. 3
1297
PHARMACOLOGY
Discussion
We used COX-2-null mice to investigate the role of COX-2 in the
mechanisms of cerebral ischemic injury. We found that, after
MCA occlusion, COX-2-null mice do not express COX-2 mRNA
and have a reduction in the elevation in PGE
2
produced by
cerebral ischemia. Furthermore, the volume of brain damage
produced by MCA occlusion is markedly reduced in COX-2-null
mice. The reduction in injury volume, like the elevation in
COX-2 mRNA and PGE
2
, is more marked in homozygous than
in heterozygous null mice. These findings provide strong evi-
dence that COX-2 reaction products are involved in ischemic
brain damage. Furthermore, the observation that, in COX-2-null
mice, infarct volume is also reduced 24 h after MCA occlusion,
supports the hypothesis that COX-2 is also involved in patho-
genic events that take place in the early stages of cerebral
ischemia.
The neuroprotection observed in COX-2-null mice cannot be
attributed to differences in the intensity of the ischemic insult,
because the reduction in CBF produced by MCA occlusion is
comparable in COX-2
and
mice. However, COX-
2-null mice were found to be more resistant to the damage
produced by the glutamate receptor antagonist NMDA. The
reduction in excitotoxicity cannot be a consequence of alter-
ations in the NMDA receptors in COX-2-null mice, because
acute administration of the COX-2 inhibitor NS398 produced
comparable attenuation in NMDA-induced injury. Conversely,
the neuroprotection exerted by NS398 is unlikely to be due to
effects of the drug unrelated to COX-2 inhibition, because
NS398 did not confer protection in COX-2-null mice. These
observations, collectively, provide evidence that COX-2 reaction
products contribute to NMDA-mediated cytotoxicity. This con-
clusion is also supported by the observations that NS398 protects
neuronal cultures from NMDA (47) and that neuronal cultures
from transgenic mice overexpressing COX-2 are more suscep-
tible to glutamate excitotoxicity (48). Considering the critical
role that NMDA receptors play in ischemic brain injury (see ref.
12 for a review), the data suggest that attenuation of glutamate
receptor-dependent ischemic damage contributes to the reduc-
tion in ischemic injury observed in COX-2-null mice.
The mechanisms by which COX-2 contributes to neurotoxicity
remain to be defined. Superoxide radicals, a COX-2 reaction
product (49, 50), are well known to participate in ischemic brain
injury and could mediate tissue damage either directly or by
reacting with nitric oxide to form the strong oxidant peroxyni-
trite (see ref. 51 for a review). Furthermore, COX-2 reaction
products could activate poly(ADP-ribose) polymerase, an en-
zyme involved in DNA repair that contributes to neurotoxicity
(52, 53). In addition, PGE
2
, a major reaction product of COX-2
(54), could mediate toxicity by facilitating glutamate release
from astrocytes (55). However, the role of PGE
2
in neurotoxicity
remains controversial because this prostaglandin has also been
reported to counter the cytotoxic effects of glutamate (56).
However, attenuation of glutamate neurotoxicity is not the
sole mechanism of the protection from ischemic injury observed
in COX-2-null mice. Cerebral ischemia is associated with an
inflammatory reaction that contributes to tissue damage (57).
COX-2 is a critical factor in the cytotoxicity associated with
inflammation (21). It is therefore likely that COX-2 plays a role
also in the mechanisms by which inflammation contributes to
ischemic damage. This hypothesis is supported by the observa-
tion that the COX-2 inhibitor NS398 reduces ischemic injury
even when administered6hafterMCAocclusion (28). At this
time after focal ischemia, activation of NMDA receptors does
not contribute to the damage, as evidenced by the fact that
NMDA receptor antagonists are no longer protective (58).
Therefore, the evidence suggests that COX-2 also plays a role in
the late stages of ischemic brain injury.
Fig. 4. Effect of NS398 on the lesion produced by direct microinjection of
NMDA into the cerebral cortex of C57BL
6J mice. (A) PGE2concentration in
the injured cortex, 3 h after NMDA injection, in mice receiving vehicle (n5)
or NS398 (n5; *,P0.05 from vehicle). (Band C) Thionin-stained brain
sections showing the lesion produced by NMDA (arrow) in mice receiving
vehicle or NS398. (D) Effect of NS398 (n6) or vehicle (n7) on the volume
of the lesion produced by NMDA (*,P0.05 from vehicle). (E) Effect of
NS398 on the lesion area at different rostrocaudal levels from the anterior
commissure.
Fig. 5. Lesion produced by microinjection of NMDA into the cerebral cortex
of COX-2
and
mice. (Aand B) Thionin-stained brain sections
showing the lesion produced by NMDA injection (arrow) in COX-2
(A)
and
(B) mice. (C) Effect of NS398 on the volume of the lesion produced
by NMDA microinjection in COX-2
and
mice (n6 per group; *,P
0.05 from vehicle). (D) Lesion area at different rostrocaudal levels relative to
the anterior commissure in COX-2
and
mice (n6 per group).
1298
www.pnas.org Iadecola et al.
The findings of the present study suggest that COX-2 is a
promising pharmacological target for the treatment of ischemic
stroke. COX-2 inhibition offers several advantages over other
prospective neuroprotective strategies. First, by targeting both
‘‘early’’ and ‘‘late’’ components of ischemic injury, COX-2
inhibitors can act as a bimodal neuroprotective strategy, with a
likelihood of success greater than that of unimodal therapies.
Second, potent and selective COX-2 inhibitors have already
proved to be relatively safe and well tolerated (22). Therefore,
examining their efficacy in stroke patients would be easier than
testing potential neuroprotective agents with an unknown safety
profile. Third, a large number of patients already take COX-2
inhibitors for treatment of osteoarthritis or pain (22). Epidemi-
ological studies could provide important clues to the usefulness
of COX-2 inhibitors in stroke prevention and improvement of
outcome. On the basis of these considerations, COX-2 inhibition
seems an attractive therapeutic strategy for stroke and other
diseases associated with glutamate excitotoxicity.
We thank Dr. Dale Cooper for help with the measurement of blood urea
nitrogen and creatinine, Dr. Carlos Manivel for pathological examina-
tion of mouse kidneys, Ms. Tracy Aber for assistance in experiments
involving RT-PCR, and Mr. Tim Murphy and Ms. Andrea Hyde for
editorial assistance. This work was supported by National Institutes of
Health Grant NS35806.
1. Jorgensen, H. S., Nakayama, H., Pedersen, P. M., Kammersgaard, L.,
Raaschou, H. O. & Olsen, T. S. (1999) Clin. Geriatr. Med. 15, 785–799.
2. Johansson, B., Norrving, B. & Lindgren, A. (2000) Stroke (Dallas) 31, 481– 486.
3. Kita, Y., Okayama, A., Ueshima, H., Wada, M., Nozaki, A., Choudhury, S. R.,
Bonita, R., Inamoto, Y. & Kasamatsu, T. (1999) Int. J. Epidemiol. 28,
1059–1065.
4. Warlow, C. P. (1998) Lancet 352, Suppl. 3, SIII1– 4.
5. Lyden, P. D., Grotta, J. C., Levine, S. R., Marler, J. R., Frankel, M. R. & Brott,
T. G. (1997) Neurology 49, 14 –20.
6. Kwiatk owski, T. G., Libman, R. B., Frankel, M., Tilley, B. C., Morgenstern,
L. B., Lu, M., Broderick, J. P., Lewandowski, C. A., Marler, J. R., Levine, S. R.
& Brott, T. (1999) N. Engl. J. Med. 340, 1781–1787.
7. Hamann, G. F., del Zoppo, G. J. & von Kummer, R. (1999) Thromb. Haemost.
82, Suppl. 1, 92–94.
8. Lyden, P. D. (1999) Prog. Cardiovasc. Dis. 42, 175–183.
9. Steiner, T. & Hacke, W. (1998) Eur. Neurol. 40, 1–8.
10. Benveniste, H., Drejer, J., Schousboe, A. & Diemer, N. H. (1984) J. Neurochem.
43, 1369–1374.
11. Butcher, S. P., Bullock, R., Graham, D. I. & McCulloch, J. (1990) Stroke
(Dallas) 21, 1727–1733.
12. Lee, J. M., Zipfel, G. J. & Choi, D. W. (1999) Nature (London) 399, A7–A14.
13. Dereski, M. O., Chopp, M., Knight, R. A., Rodolosi, L. C. & Garcia, J. H.
(1993) Acta. Neuropathol. 85, 327–333.
14. Du, C., Hu, R., Csernansky, C. A., Hsu, C. Y. & Choi, D. W. (1996) J. Cereb.
Blood Flow Metab. 16, 195–201.
15. Endres, M., Namura, S., Shimizu-Sasamata, M., Waeber, C., Zhang, L.,
Gomez-Isla, T., Hyman, B. T. & Moskowitz, M. A. (1998) J. Cereb. Blood Flow
Metab. 18, 238–247.
16. Chopp, M., Li, Y., Jiang, N., Zhang, R. L. & Prostak, J. (1996) J. Cereb. Blood
Flow Metab. 16, 578–584.
17. Connolly, E. S., Jr., Winfree, C. J., Springer, T. A., Naka, Y., Liao, H., Yan,
S. D., Stern, D. M., Solomon, R. A., Gutierrez-Ramos, J. C. & Pinsky, D. J.
(1996) J. Clin. Invest. 97, 209–216.
18. Dirnagl, U., Iadecola, C. & Moskowitz, M. A. (1999) Trends Neurosci. 22,
391–397.
19. Vane, J. R., Bakhle, Y. S. & Botting, R. M. (1998) Annu. Rev. Pharmacol.
Toxicol. 38, 97–120.
20. Seibert, K., Zhang, Y., Leahy, K., Hauser, S., Masferrer, J. & Isakson, P. (1997)
Adv. Exp. Med. Biol. 400A, 167–170.
21. Seibert, K., Masferrer, J., Zhang, Y., Gregory, S., Olson, G., Hauser, S., Leahy,
K., Perkins, W. & Isakson, P. (1995) Agents Actions Suppl. 46, 41–50.
22. Hawkey, C. J. (1999) Lancet 353, 307–314.
23. Willoughby, D. A., Moore, A. R. & Colville-Nash, P. R. (2000) Lancet 355,
646– 648.
24. Yamagata, K., Andreasson, K. I., Kaufmann, W. E., Barnes, C. A. & Worley,
P. F. (1993) Neuron 11, 371–386.
25. Kaufmann, W. E., Worley, P. F., Pegg, J., Bremer, M. & Isakson, P. (1996) Proc.
Natl. Acad. Sci. USA 93, 2317–2321.
26. Breder, C. D., Dewitt, D. & Kraig, R. P. (1995) J. Comp. Neurol. 355, 296–315.
27. Miettinen, S., Fusco, F. R., Yrjanheikki, J., Keinanen, R., Hirvonen, T.,
Roivainen, R., Narhi, M., Hokfelt, T. & Koistinaho, J. (1997) Proc. Natl. Acad.
Sci. USA 94, 6500–6505.
28. Nogawa, S., Zhang, F., Ross, M. E. & Iadecola, C. (1997) J. Neurosci. 17,
2746–2755.
29. Pasinetti, G. M. (1998) J. Neurosci. Res. 54, 1–6.
30. Iadecola, C., Forster, C., Nogawa, S., Clark, H. B. & Ross, M. E. (1999) Acta
Neuropathol. 98, 9–14.
31. Hara, K., Kong, D. L., Sharp, F. R. & Weinstein, P. R. (1998) Neurosci. Lett.
256, 53–56.
32. Cole, D. J., Patel, P. M., Reynolds, L., Drummond, J. C. & Marcanton io, S.
(1993) J. Pharmacol. Exp. Ther. 266, 1713–1717.
33. Lehmann, J. M., Lenhard, J. M., Oliver, B. B., Ringold, G. M. & Kliewer, S. A.
(1997) J. Biol. Chem. 272, 3406–3410.
34. Grilli, M., Pizzi, M., Memo, M. & Spano, P. (1996) Science 274, 1383–1385.
35. Morham, S. G., Langenbach, R., Loftin, C. D., Tiano, H. F., Vouloumanos, N.,
Jennette, J. C., Mahler, J. F., Kluckman, K. D., Ledford, A., Lee, C. A., et al.
(1995) Cell 83, 473–482.
36. Niwa, K., Araki, E., Morham, S. G., Ross, M. E. & Iadecola, C. (2000)
J. Neurosci. 20, 763–770.
37. Langenbach, R., Loftin, C. D., Lee, C. & Tiano, H. (1999) Ann. N.Y. Acad. Sci.
889, 52–61.
38. Iadecola, C., Zhang, F., Casey, R., Nagayama, M. & Ross, M. E. (1997)
J. Neurosci. 17, 9157–9164.
39. Nogawa, S., Forster, C., Zhang, F., Nagayama, M., Ross, M. E. & Iadecola, C.
(1998) Proc. Natl. Acad. Sci. USA 95, 10966–10971.
40. Chomczynski, P. & Sacchi, N. (1987) Anal. Biochem. 162, 156 –159.
41. Iadecola, C., Zhang, F., Casey, R., Clark, H. B. & Ross, M. E. (1996) Stroke
(Dallas) 27, 1373–1380.
42. Galea, E. & Feinstein, D. L. (1992) PCR Methods Appl. 2, 66– 69.
43. Diviacco, S., Norio, P., Sentilin, L., Menzo, S., Clementi, M., Biamonti, G.,
Riva, S., Falaschi, A. & Giacca, M. (1992) Gene 122, 313–320.
44. Lin, T.-N., He, Y. Y., Wu, G., K han, M. & Hsu, C. Y. (1993) Stroke (Dallas)
24, 117–121.
45. Loeb, W. F. & Quimby, F. W. (1989) The Clinical Chemistry of Laborator y
Animals (Pergamon, New York).
46. Hossmann, K.-A. (1994) Ann. Neurol. 36, 557–565.
47. Hewett, S. J., Uliasz, T. F., Vidwans, A. S. & Hewett, J. A. (2000) J. Pharmacol.
Exp. Ther. 293, 417–425.
48. Kelley, K. A., Ho, L., Winger, D., Freire-Moar, J., Borelli, C. B., Aisen, P. S.
& Pasinetti, G. M. (1999) Am. J. Pathol. 155, 995–1004.
49. Kontos, H. A., Wei, E. P., Povlishock, J. T., Dietrich, W. D., Magiera, C. J. &
Ellis, E. F. (1980) Science 209, 1242–1245.
50. Chan, P. H. & Fishman, R. A. (1980) J. Neurochem. 35, 1004 –1007.
51. Chan, P. H. (1996) Stroke (Dallas) 27, 1124 –1129.
52. Eliasson, M. J., Sampei, K., Mandir, A. S., Hurn, P. D., Traystman, R. J., Bao,
J., Pieper, A., Wang, Z. Q., Dawson, T. M., Snyder, S. H. & Dawson, V. L.
(1997) Nat. Med. 3, 1089 –1095.
53. Endres, M., Scott, G., Namura, S., Salzman, A. L., Huang, P. L., Moskowitz,
M. A. & Szabo, C. (1998) Neurosci. Lett. 248, 41– 44.
54. Brock, T. G., McNish, R. W. & Peters-Golden, M. (1999) J. Biol. Chem. 274,
11660–11666.
55. Bezzi, P., Carmignoto, G., Pasti, L., Vesce, S., Rossi, D., Rizzini, B. L., Pozzan,
T. & Volterra, A. (1998) Nature (London) 391, 281–285.
56. Cazevieille, C., Muller, A., Meynier, F., Dutrait, N. & Bonne, C. (1994)
Neurochem. Int. 24, 395–398.
57. Barone, F. C. & Feuerstein, G. Z. (1999) J. Cereb. Blood Flow Metab. 19,
819– 834.
58. Hossmann, K.-A. (1994) Brain Pathol. 4, 23–36.
Iadecola et al. PNAS
January 30, 2001
vol. 98
no. 3
1299
PHARMACOLOGY
... Thus, COX-2 is an attractive target for stroke therapy [36]. In cerebral ischemia, COX-2 contributes to glutamate excitotoxicity [37,38], a vital factor in the Ca 2þ dysregulation initiating the ischemic cascade [39]. ...
Article
Full-text available
Background Stroke is the leading cause of mortality and morbidity worldwide, and an effective therapeutic strategy for the prevention of patients with cerebral ischemia induced brain injury is lacking. Traditional Chinese medicine with neuroprotective activities might be beneficial and provide alternative therapeutic opportunities for cerebral ischemia. Purposes This study aimed to evaluate the neuroprotection and possible mechanisms of Gueichih-Fuling-Wan (GFW), its’ constitutive herbs, and their active compounds on cerebral ischemia/reperfusion (I/R)-induced brain injury in rodents. Methods Various doses of extracts (0.25, 0.5, and 1.0 g/kg) of GFW and five constituent herbs (Cinnamomi Cortex, CC; Poria cocos, PC; Paeonia lactifloa, PL; Paeonia suffruticosa, PS and Prunus perisica, PP) were orally administered. Different doses of active compounds (0.5, 1.0, and 2.0 mg/kg) of GFW such as cinnamaldehyde, cinnamic acid (from CC), paeoniflorin (from PL), and paeonol (from PS) were intraperitoneally administered. Their effects on cerebral ischemia/ reperfusion (I/R)induced brain injury in rodents were evaluated. Results GFW, its’ constituent herbs, and the active compounds reduced the infarct area dose-dependently (***P < 0.001). Cinnamaldehyde showed the most significant reduction (***P < 0.001). Therefore, trans-cinnamaldehyde (TCA) was further used to evaluate the neuroprotective mechanism of the I/R-induced brain injury. TCA (10, 20, 30 mg/ kg, p.o.) showed an inhibitory effect of I/R-induced brain damage in mice in a dose-dependent manner. Besides, GFW and TCA dose-dependently reduced the COX-2 protein expression level, and TCA reduced the TUNEL (+) apoptosis. TCA dose-dependently increased the pro-survival NR2A and Bcl-2 protein expression level and decreased the pro-apoptotic NR2B and cytochrome c, caspase 9, and caspase 3 expression (***P < 0.001). Conclusion The above data revealed that GFW, its’ constituent herbs, and active compounds protected against I/R-induced brain injury in rodents. TCA from CC might participate in GFW protecting against cerebral ischemia-induced brain injury by inhibiting neuroinflammation and apoptosis.
... In the brain, COX-2 is expressed in discrete populations of neurons, enriched in the cortex and hippocampus [44], and has been implicated to brain functions and neurological diseases, including stroke, epilepsy, and Alzheimer's disease [45][46][47][48]. Numerous studies have shown an elevated expression level of COX-2 in the brains of individuals and animals with epilepsy [49,50,44,51,37,[52][53][54]. ...
Article
Full-text available
Arachidonic acid (AA), an important polyunsaturated fatty acid in the brain, is hydrolyzed by a direct action of phospholipase A2 (PLA2) or through the combined action of phospholipase C and diacylglycerol lipase, and released into the cytoplasm. Various derivatives of AA can be synthesized mainly through the cyclooxygenase (COX), lipoxygenase (LOX) and cytochrome P450 (P450) enzyme pathways. AA and its metabolic enzymes and metabolites play important roles in a variety of neurophysiological activities. The abnormal metabolites and their catalytic enzymes in the AA cascade are related to the pathogenesis of various central nervous system (CNS) diseases, including epilepsy. Here, we systematically reviewed literatures in PubMed about the latest randomized controlled trials, animal studies and clinical studies concerning the known features of AA, its metabolic enzymes and metabolites, and their roles in epilepsy. The exclusion criteria include non-original studies and articles not in English.
... During cerebral ischemia, when inflammatory cells infiltrate the injured brain, www.nature.com/scientificreports/ Ptgs2 is found to be upregulated 38 . Furthermore, transient cerebral ischemia has been shown to induce marked apoptosis and inflammatory responses such as NF-κB activation 39 . ...
Article
Full-text available
Ischemic stroke (IS) is associated with changes in gene expression patterns in the ischemic penumbra and extensive neurovascular inflammation. However, the key molecules related to the inflammatory response in the acute phase of IS remain unclear. To address this knowledge gap, conducted a study using Gene Set Enrichment Analysis (GSEA) on two gene expression profiles, GSE58720 and GSE202659, downloaded from the GEO database. We screened differentially expressed genes (DEGs) using GEO2R and analyzed 170 differentially expressed intersection genes for Kyoto Encyclopedia of Genes and Genomes (KEGG) pathway enrichment and Gene Ontology (GO) analysis. We also used Metascape, DAVID, STRING, Cytoscape, and TargetScan to identify candidate miRNAs and genes. The targeted genes and miRNA molecule were clarified using the mice middle cerebral artery occlusion-reperfusion (MCAO/R) model. Our findings revealed that 170 genes were correlated with cytokine production and inflammatory cell activation, as determined by GO and KEGG analyses. Cluster analysis identified 11 hub genes highly associated with neuroinflammation: Ccl7, Tnf, Ccl4, Timp1, Ccl3, Ccr1, Sele, Ccr2, Tlr4, Ptgs2, and Il6. TargetScan results suggested that Ptgs2, Tlr4, and Ccr2 might be regulated by miR-202-3p. In the MCAO/R model, the level of miR-202-3p decreased, while the levels of Ptgs2, Tlr4, and Ccr2 increased compared to the sham group. Knockdown of miR-202-3p exacerbated ischemic reperfusion injury (IRI) through neuroinflammation both in vivo and in vitro. Our study also demonstrated that mRNA and protein levels of Ptgs2, Tlr4, and Ccr2 increased in the MCAO/R model with miR-202-3p knockdown. These findings suggest that differentially expressed genes, including Ptgs2, Tlr4, and Ccr2 may play crucial roles in the neuroinflammation of IS, and their expression may be negatively regulated by miR-202-3p. Our study provides new insights into the regulation of neuroinflammation in IS.
... COX inhibitors, which are effective anti-inflammatory drugs, have been used after transient global cerebral ischemia in rodents to improve the delayed death of hippocampal CA1 neurons [21]. In addition, COX-2 gene expression-deficient mice exhibited a protective effect in terms of reducing brain damage in a middle cerebral artery occlusion experiment, indicating that inhibiting COX-2 may be an important factor in the reduction of glutamate neurotoxicity [22]. Indeed, previous studies reported COX-2 to be the upstream factor of glutamate excitotoxicity, affecting cell fate by regulating AMPA glutamate receptors [23]. ...
Article
Full-text available
The study aimed to identify TUG1 as an essential regulator of apoptosis in HT22 (mouse hippocampal neuronal cells) by direct interaction with the RNA-binding protein HuR. In order to study the role of TUG1 in the context of ischemia, we used mouse hippocampal neuronal cells treated with oxyglucose deprivation to establish an in-vitro ischemia model. A bioinformatic analysis and formaldehyde RNA immunoprecipitation (fRIP) were used to investigate the biological functions. A Western blot assay and reverse transcription polymerase chain reaction were used to explore the expression of the molecules involved. A cell proliferation and cytotoxicity assay was performed to detect neuronal apoptosis. TUG1 exhibits a localization-specific expression pattern in HT22 cells under OGD treatment. The bioinformatics analysis showed a strong correlation between the TUG1 and HuR as predicted, and this interaction was subsequently confirmed by fRIP-qPCR. We found that HuR was translocated from the nucleus to the cytoplasm after ischemia treatment and subsequently targeted and stabilized COX-2 mRNA, which led to elevated COX-2 mRNA levels and apoptosis of the HT22 cells. Furthermore, nuclear-specific disruption of TUG1 prevented the translocation of HuR to the cytoplasm and decreased COX-2 mRNA expression, resulting in increased cell viability and partially reversed apoptosis. In conclusion, it was demonstrated that TUG1 accelerates the process of apoptosis by promoting the transfer of HuR to the cytoplasm and stabilizing COX-2 mRNA. These results provide useful information concerning a therapeutic target for ischemic stroke.
... Les inhibiteurs de la COX-2 dont le NS-398 et le nimésulide réduisent l'infarctus (Iadecola et al., 2001b;Sugimoto and Iadecola, 2003). Le NS-398 et le nimésulide sont toujours efficaces même lorsque leur administration est effectuée jusqu'à 24 heures après l'ischémie. ...
Thesis
Les accidents vasculaires cérébraux (AVC) constituent la 2ème cause de mortalité dans le monde et la 1ère chez les femmes en France. Pour les AVC ischémiques (AVCi), seules des stratégies de recanalisation pharmacologique ou mécanique ont été approuvées mais aucune stratégie protectrice n'est aujourd'hui disponible. Bien que le rôle délétère du stress oxydant ait été clairement établi dans les lésions neuronales et vasculaires à la suite d'une ischémie cérébrale (IC) dans les études précliniques, aucune stratégie anti-oxydante n'a démontré d'efficacité clinique à ce jour. Or, les nanoparticules d'oxyde de cérium (NPC) possèdent de multiples capacités antioxydantes (enzymatique et non enzymatique). Afin d'améliorer la biocompatibilité des NPC, la société Specific Polymers® a développé des copolymères de polyéthylène glycol (PEG)/ polyméthacrylate de méthyle/ phosphonate pour recouvrir leur surface. De plus, ces polymères peuvent être fonctionnalisés avec un peptide ciblant l'endothélium ce qui permettrait d'y concentrer l'effet antioxydant des NPC, afin de réduire la survenue d'hémorragies cérébrales, complications graves chez les patients victimes d'AVCi. L'objectif de cette thèse est d'évaluer l'impact du recouvrement des NPC sur leur potentiel thérapeutique dans l'IC. Les études ont été menées in vitro pour établir la toxicité, l'effet antioxydant et l'internalisation cellulaire des NPC et in vivo, pour examiner leur biodistribution et leur toxicité, ainsi que leur potentiel thérapeutique dans un modèle d'IC. Les études in vitro ont été effectuées sur des cellules endothéliales cérébrales murines de la lignée b.End3. Nous avons démontré que les NPC n'induisaient ni mortalité, ni perturbation de l'activité métabolique jusqu'à 100µg/ml. A 1000µg/mL, les NPC nues augmentent la mortalité, contrairement aux NPC PEGylées. Nous avons modélisé l'excitotoxicité survenant lors d'une IC et qui contribue au stress oxydant, grâce à un traitement des cellules par le glutamate. L'augmentation de la production d'espèces réactives de l'oxygène par les cellules b.End3 et l'oxydation des acides nucléiques dans ces conditions ont été réduites par les NPC, démontrant que leur recouvrement n'interfère pas avec leurs propriétés anti-oxydantes. La fonctionnalisation des NPC a permis le greffage d'un fluorophore pour suivre leur internalisation par cytométrie en flux et microscopie confocale. Nous avons ainsi mis en évidence que les NPC étaient rapidement internalisées dans les cellules b.End3. Des études de microscopie électronique ont ensuite montré que les NPC sont principalement localisées dans des endosomes périnucléaires. Enfin, nous avons réalisé le greffage sur les NPC d'un peptide ciblant une protéine d'adhésion vasculaire surexprimée lors de l'IC. La suite de ces études consistera à vérifier l'interaction spécifique de ces NPC avec la molécule d'adhésion. Les études in vivo ont permis d'établir la biodistribution des NPC chez des souris Swiss : des NPC sont retrouvées durant les premières heures suivant leur injection, avant leur élimination par voie rénale. L'histopathologie n'a révélé aucune toxicité des NPC sur le foie, les reins, la rate, les poumons et le cerveau de ces souris et aucune modification de leur numération sanguine n'a été observée. Les NPC ont ensuite été administrées dans un modèle murin d'IC, mais n'ont pas réduit le volume de la lésion dans nos conditions. En conclusion, le recouvrement des NPC par des polymères innovants a réduit leur toxicité sans altérer leurs capacités antioxydantes et leur internalisation dans des cellules endothéliales cérébrales. L'absence d'accumulation à long terme et de toxicité in vivo sont encourageantes quant à leur biocompatibilité. Bien que les NPC n'aient pas montré d'effet protecteur in vivo, celles ciblant l'endothélium pourraient réduire les lésions vasculaires et le risque hémorragique consécutif à une IC.
... Upregulation of the p38 MAPK pathway results in overexpression of inflammatory mediators such as i-NOS and COX-2, which are also related to oxidative stress [14,15,38]. Oxidative stress mediated by i-NOS/COX-2 increases the production of pro-inflammatory cytokines, including IL-2, IL-6, and tumor necrosis factor-α through the upregulation of nuclear factor kappa-B signaling [39][40][41]. In this study, POP significantly suppressed MCAO/R-mediated upregulation of p-p38 MAPK expression in the brain, and the increased expression of i-NOS and COX-2 after MCAO/R was prevented by POP treatment. ...
Article
Full-text available
1,3-Dipalmitoyl-2-oleoylglycerol (POP) is a triacylglyceride found in oils from various natural sources, including palm kernels, sunflower seeds, and rice bran. In the current study, the neuroprotective effects and the specific mechanism of POP derived from rice bran oil were investigated for the first time using the middle cerebral artery occlusion/reperfusion (MCAO/R) model in rats. Orally administered POP at 1, 3, or 5 mg/kg (three times: 0.5 h before MCAO, after 1 h of MCAO, and after 1 h of reperfusion) markedly reduced the MCAO/R-induced infarct/edema volume and neurobehavioral deficits. Glutathione depletion and the oxidative degradation of lipids in the rat brain induced by MCAO/R were prevented by POP administration. The upregulation of phosphorylated p38 MAPKs, inflammatory factors (inducible nitric oxide synthase (i-NOS) and cyclooxygenase-2 (COX-2)), and pro-apoptotic proteins (B-cell lymphoma-2 (Bcl-2) associated X protein (Bax) and cleaved caspase-3) and the downregulation of the anti-apoptotic protein (Bcl-2) in the ischemic brain were significantly inhibited by POP administration. In addition, downregulation of phosphatidylinositol 3′-kinase (PI3K), phosphorylated protein kinase B (Akt), and phosphorylated cyclic (adenosine monophosphate) AMP responsive element-binding protein (CREB) expression in the ischemic brain was inhibited by POP administration. These results suggest that POP might exert neuroprotective effects by inhibition of p38 MAPK and activation of PI3K/Akt/CREB pathway, which is associated with anti-oxidant, anti-apoptotic, and anti-inflammatory action. From the above results, the present study provides evidence that POP might be effectively applied for the management of cerebral ischemia-related diseases.
... The generation of ROS by cyclooxygenases (COXs) and lipoxygenases (LOXs) have also been targeted in animal models of stroke. COX-2 knockout in mice and COX-2 inhibition with NS-398 resulted in reduced infarct size after middle cerebral artery occlusion [140,141]. Similarly, 12/15-LOX knockout or inhibition with baicalein or LOXBlock-1 led to reduced infarct size after transient middle cerebral artery occlusion in animal experiments [142,143]. ...
Article
Full-text available
Recanalization therapy is increasingly used in the treatment of acute ischemic stroke. However, in about one third of these patients, recanalization is followed by ischemia/reperfusion injuries, and clinically to worsening of the neurological status. Much research has focused on unraveling the involved mechanisms in order to prevent or efficiently treat these injuries. What we know so far is that oxidative stress and mitochondrial dysfunction are significantly involved in the pathogenesis of ischemia/reperfusion injury. However, despite promising results obtained in experimental research, clinical studies trying to interfere with the oxidative pathways have mostly failed. The current article discusses the main mechanisms leading to ischemia/reperfusion injuries, such as mitochondrial dysfunction, excitotoxicity, and oxidative stress, and reviews the clinical trials with antioxidant molecules highlighting recent developments and future strategies.
Article
Background Many medical experts prescribe indomethacin because of its anti‐inflammatory, analgesic, tocolytic, and duct closure effects. This article presents an evaluation of the enduring impact of indomethacin on neonatal rats with hypoxic–ischemic (HI) insults, employing behavioral tests as a method of assessment. Methods The experiment was conducted on male Wistar‐Albino rats weighing 10 to 15 g, aged between seven and 10 days. The rats were divided into three groups using a random allocation method as follows: hypoxic ischemic encephalopathy (HIE) group, HIE treated with indomethacin group (INDO), and Sham group. A left common carotid artery ligation and hypoxia model was applied in both the HIE and INDO groups. The INDO group was treated with 4 mg/kg intraperitoneal indomethacin every 24 h for 3 days, while the Sham and HIE groups were given dimethylsulfoxide (DMSO). After 72 h, five rats from each group were sacrificed and brain tissue samples were stained with 2,3,5‐Triphenyltetrazolium chloride (TCC) for infarct‐volume measurement. Seven rats from each group were taken to the behavioral laboratory in the sixth postnatal week (PND42) and six from each group were sacrificed for the Evans blue (EB) experiment for blood–brain barrier (BBB) integrity evaluation. The open field (OF) test and Morris water maze (MWM) tests were performed. After behavioral tests, brain tissue were obtained and stained with TCC to assess the infarct volume. Results The significant increase in the time spent in the central area and the frequency of crossing to the center in the INDO group compared with the HIE group indicated that indomethacin decreased anxiety‐like behavior ( p < 0.001, p < 0.05). However, the MWM test revealed that indomethacin did not positively affect learning and memory performance ( p > 0.05). Additionally, indomethacin significantly reduced infarct volume and neuropathological grading in adolescence ( p < 0.05), although not statistically significant in the early period. Moreover, the EB experiment demonstrated that indomethacin effectively increased BBB integrity ( p < 0.05). Conclusions In this study, we have shown for the first time that indomethacin treatment can reduce levels of anxiety‐like behavior and enhance levels of exploratory behavior in a neonatal rat model with HIE. It is necessary to determine whether nonsteroidal anti‐inflammatory agents, such as indomethacin, should be used for adjuvant therapy in newborns with HIE.
Article
Full-text available
Background: A. chinense frequently used in Miao medicine to treat rheumatic diseases. However, as a famous toxic herb, Alangium chinense and its representative components exhibit ineluctable neurotoxicity, thus creating significant challenges for clinical application. The combined application with compatible herbs in Jin-Gu-Lian formula attenuates such neurotoxicity according to the compatible principle of traditional Chinese medicines. Purpose: We aimed to investigate the detoxification of the compatible herbs in Jin-Gu-Lian formula on A. chinense-induced neurotoxicity and investigate its mechanism. Methods: Neurobehavioral and pathohistological analysis were used to determine the neurotoxicity in rats administered with A. chinense extract (AC), extract of compatible herbs in Jin-Gu-Lian formula (CH) and combination of AC with CH for 14 days. The mechanism underlying the reduction of toxicity by combination with CH was assessed by enzyme-linked immunosorbent assays, spectrophotometric assays, liquid chromatography tandem-mass spectrometry and real-time reverse transcription-quantitative polymerase chain reaction. Results: Compatible herbs attenuated the AC-induced neurotoxicity as evidenced by increased locomotor activity, enhanced grip strength, the decreased frequency of AC-induced morphological damage in neurons, as well as a reduction of neuron-specific enolase (NSE) and neurofilament light chain (NEFL) levels. The combination of AC and CH ameliorated AC-induced oxidative damage by modulating the activities of superoxide dismutase (SOD) and glutathione peroxidase (GSH-Px), and total antioxidant capacity (T-AOC). AC treatment significantly reduced the levels of monoamine and acetylcholine neurotransmitters in the brains of rats, including acetylcholine (Ach), dopamine (DA), 3,4-dihydroxyphenylacetic acid (DOPAC), homovanillic acid (HVA), norepinephrine (NE), and serotonin (5-HT). Combined AC and CH treatment regulated the abnormal concentrations and metabolisms of neurotransmitters. Pharmacokinetic studies showed that the co-administration of AC and CH significantly decreased plasma exposure levels of two main components of AC, as evidenced by the reduction of maximum plasma concentration (Cmax), area under the plasma concentration-time curve (AUC) compared to AC. In addition, the AC-induced downregulation in mRNA expression of cytochrome P450 enzymes was significantly reduced in response to combined AC and CH treatment. Conclusion: Compatible herbs in Jin-Gu-Lian formula alleviated the neurotoxicity induced by A. chinense by ameliorating oxidative damage, preventing abnormality of neurotransmitters and modulating pharmacokinetics.
Article
Objective To investigate whether tert-Butylhydroquinone (TBHQ) can ameliorate oxidative stress and inflammation induced by glutamate excitotoxicity, and mediate retinal ganglion cell (RGC) damage by activating the nuclear factor erythroid 2-related factor 2 (Nrf2) signaling pathway and inhibiting the nuclear factor kappa B (NF-κB) signaling pathway. Materials and methods TBHQ was used to treat a glutamate excitotoxicity model of retinal cell line 28 and C57 mice. Damage to RGCs and visual function were assessed using flash visual evoked potential (FVEP), immunofluorescence, propidium iodide staining, and hematoxylin and eosin staining. Knockdown of Nrf2 used Nrf2 shRNA. The expression levels of related proteins were detected using western blot and immunofluorescence. Results Glutamate excitotoxicity down-regulated Nrf2 expression in vitro and in vivo. Nuclear factor erythroid2-related factor 2 activation by TBHQ reduced the damage to retinal ganglion cells, reduced the thinning of the whole retina and the ganglion cell complex, and shortened the latency of the FVEP forward wave after injury. In addition, the levels of NAD(P)H quinone dehydrogenase 1 (NQO1), heme oxygenase 1 (HO-1), and Nrf2 increased significantly, and those of cyclooxygenase-2 (COX2) and NF-κB decreased significantly, after TBHQ treatment. Compared with TBHQ treatment group, the expression level of p-p65 in shRNA transfected group was increased, but still lower than that in Glu group. Conclusion The protective effect of TBHQ on RGC loss under glutamate excitotoxicity might be related to the activation of the Nrf2 signaling pathway, anti-oxidative stress, inhibition of NF-κB activation, and inhibition of retinal inflammation. Thus, TBHQ might be used to treat glutamate excitotoxicity -related retinopathy.
Article
Full-text available
The prostanoid-synthesizing enzyme cyclooxygenase-2 (COX-2) is expressed in selected cerebral cortical neurons and is involved in synaptic signaling. We sought to determine whether COX-2 participates in the increase in cerebral blood flow produced by synaptic activity in the somatosensory cortex. In anesthetized mice, the vibrissae were stimulated mechanically, and cerebral blood flow was recorded in the contralateral somatosensory cortex by a laser-Doppler probe. We found that the COX-2 inhibitor NS-398 attenuates the increase in somatosensory cortex blood flow produced by vibrissal stimulation. Furthermore, the flow response was impaired in mice lacking the COX-2 gene, whereas the associated increase in whisker-barrel cortex glucose use was not affected. The increases in cerebral blood flow produced by hypercapnia, acetylcholine, or bradykinin were not attenuated by NS-398, nor did they differ between wild-type and COX-2 null mice. The findings provide evidence for a previously unrecognized role of COX-2 in the mechanisms coupling synaptic activity to neocortical blood flow and provide an insight into one of the functions of constitutive COX-2 in the CNS.
Article
Full-text available
A new method of total RNA isolation by a single extraction with an acid guanidinium thiocyanate-phenol-chloroform mixture is described. The method provides a pure preparation of undegraded RNA in high yield and can be completed within 4 h. It is particularly useful for processing large numbers of samples and for isolation of RNA from minute quantities of cells or tissue samples.
Article
Full-text available
Repetitive spreading depression (SD) waves, involving depolarization of neurons and astrocytes and up-regulation of glucose consumption, is thought to lower the threshold of neuronal death during and immediately after ischemia. Using rat models for SD and focal ischemia we investigated the expression of cyclooxygenase-1 (COX-1), the constitutive form, and cyclooxygenase-2 (COX-2), the inducible form of a key enzyme in prostaglandin biosynthesis and the target enzymes for nonsteroidal anti-inflammatory drugs. Whereas COX-1 mRNA levels were undetectable and uninducible, COX-2 mRNA and protein levels were rapidly increased in the cortex, especially in layers 2 and 3 after SD and transient focal ischemia. The cortical induction was reduced by MK-801, an N-methyl-d-aspartic acid-receptor antagonist, and by dexamethasone and quinacrine, phospholipase A2 (PLA2) inhibiting compounds. MK-801 acted by blocking SD whereas treatment with PLA2 inhibitors preserved the wave propagation. NBQX, an α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid/kainate-receptor antagonist, did not affect the SD-induced COX-2 expression, whereas COX-inhibitors indomethacin and diclofenac, as well as a NO synthase-inhibitor, NG-nitro-l-arginine methyl ester, tended to enhance the COX-2 mRNA expression. In addition, ischemia induced COX-2 expression in the hippocampal and perifocal striatal neurons and in endothelial cells. Thus, COX-2 is transiently induced after SD and focal ischemia by activation of N-methyl-d-aspartic acid-receptors and PLA2, most prominently in cortical neurons that are at a high risk to die after focal brain ischemia.
Article
Full-text available
Focal cerebral ischemia is associated with expression of both inducible nitric oxide synthase (iNOS) and cyclooxygenase-2 (COX-2), enzymes whose reaction products contribute to the evolution of ischemic brain injury. We tested the hypothesis that, after cerebral ischemia, nitric oxide (NO) produced by iNOS enhances COX-2 activity, thereby increasing the toxic potential of this enzyme. Cerebral ischemia was produced by middle cerebral artery occlusion in rats or mice. Twenty-four hours after ischemia in rats, iNOS-immunoreactive neutrophils were observed in close proximity (<20 μm) to COX-2-positive cells at the periphery of the infarct. In the olfactory bulb, only COX-2 positive cells were observed. Cerebral ischemia increased the concentration of the COX-2 reaction product prostaglandin E2 (PGE2) in the ischemic area and in the ipsilateral olfactory bulb. The iNOS inhibitor aminoguanidine reduced PGE2 concentration in the infarct, where both iNOS and COX-2 were expressed, but not in the olfactory bulb, where only COX-2 was expressed. Postischemic PGE2 accumulation was reduced significantly in iNOS null mice compared with wild-type controls (C57BL/6 or SV129). The data provide evidence that NO produced by iNOS influences COX-2 activity after focal cerebral ischemia. Pro-inflammatory prostanoids and reactive oxygen species produced by COX-2 may be a previously unrecognized factor by which NO contributes to ischemic brain injury. The pathogenic effect of the interaction between NO, or a derived specie, and COX-2 is likely to play a role also in other brain diseases associated with inflammation.
Article
In this study we describe the generation of a transgenic mouse model with neuronal overexpression of the human cyclooxygenase-2, h(COX)-2, to explore its role in excitotoxicity. We report that overexpression of neuronal hCOX-2 potentiates the intensity and lethality of kainic acid excitotoxicity in coincidence with potentiation of expression of the immediate early genes c-fos and zif-268. In vitro studies extended the in vivo findings and revealed that glutamate excitotoxicity is potentiated in primary cortico-hippocampal neurons derived from hCOX-2 transgenic mice, possibly through potentiation of mitochondrial impairment. This study is the first to demonstrate a cause-effect relationship between neuronal COX-2 expression and excitotoxicity. This model system will allow the systematic examination of the role of COX-2 in mechanisms of neurodegeneration that involve excitatory amino acid pathways.
Article
ABSTRACT Cyclooxygenase (COX)-1- and COX-2-deficient mice have unique physiological differences that have allowed investigation into the individual biological roles of the COX isoforms. In the following, the phenotypes of the two COX knockout mice are summarized, and recent studies to investigate the effects of COX deficiency on inflammatory responses and cancer susceptibility are discussed. The data suggest that both isoforms have important roles in the maintenance of physiological homeostasis and that such designations as housekeeping and/or response gene may not be entirely accurate. Furthermore, data from COX-deficient mice indicate that both isoforms can contribute to the inflammatory response and that both isoforms have significant roles in carcino-genesis.
Article
Many epidemiological studies suggest that use of non-steroidal anti-inflammatory drugs (NSAIDs) delay or slow the clinical expression of Alzheimer's disease (AD). While it has been demonstrated that neurodegeneration in AD is accompanied by specific inflammatory mechanisms, including activation of the complement cascade and the accumulation and activation of microglia, the mechanism by which NSAIDs might affect these or other pathophysiological processes relevant to AD has been unclear. New evidence that cyclooxygenase (COX) is involved in neurodegeneration along with the development of selective COX inhibitors has led to renewed interest in the therapeutic potential of NSAIDs in AD. J. Neurosci. Res. 54:1–6, 1998. © 1998 Wiley-Liss, Inc.
Article
The prostaglandin synthesizing enzyme cyclooxygenase-2 (COX-2) is up-regulated in the brain of rodents during cerebral ischemia and contributes to ischemic brain injury. This study sought to determine whether COX-2 is also up-regulated in the human brain in the acute stages of cerebral ischemic infarction. Paraffin-embedded sections from patients who died 1–2 days following infarction in the middle cerebral artery territory were processed for COX-2 immunohistochemistry. COX-2 immunoreactivity was observed in infiltrating neutrophils, in vascular cells and in neurons located at the border of the infarct. The data suggest that COX-2 up-regulation is also relevant to cerebral ischemia in humans and raise the possibility that COX-2 reaction products participate in the mechanisms of ischemic injury also in the human brain.
Article
Non-steroidal anti-inflammatory drugs (NSAIDs) are widely used in the treatment of a number of inflammatory diseases and are believed to act via inhibition of the enzyme cyclooxygenase (COX). This enzyme catalyzes the conversion of arachidonic acid to the prostaglandins (PGs). Although commercially available NSAIDs are efficacious anti-inflammatory agents, significant side effects limit their use. Recently two forms of COX were identified-a constitutively expressed COX-1 and a cytokine-inducible COX-2. Potent anti-inflammatory agents like the glucocorticoids are known to inhibit specifically the expression of COX-2 while commercially available NSAIDs like indomethacin inhibit both COX-1 and COX-2. These findings have led to the hypothesis that toxicities associated with NSAID therapy are due to inhibition of the non-regulated or constitutive form of COX (COX-1), whereas therapeutic benefit derives from inhibition of the inducible enzyme, COX-2. We have examined the relative distribution of COX-1 and COX-2 in both normal and inflamed tissues and report that COX-1 expression dominates normal tissues while COX-2 mRNA is induced at the inflammatory site. Furthermore, compounds that selectively inhibit COX-2 are anti-inflammatory without gastric toxicity.