ArticlePDF Available

Interleukin-1ß mRNA expression in ischemic rat cortex

Authors:
  • BioHighland Consulting

Abstract

Background and Pur pose: Interleukin-1ß is a proinftammatory cytokine produced by blood-borne and resident brain inftammatory cells. The present study was conducted to determine if interleukin-1ß mRNA was produced in the brain of rats subjected to permanent focal ischemia. Methods: Rat interleukin-1ß cDNA, synthesized from stimulated rat peritoneal macrophage RNA by reverse transcription and polymerase chain reaction and c10ned in plasmid Bluescript KS+, was used to evaluate the expression of interleukin-1ß mRNA in cerebral cortex from spontaneously hypertensive rats and normotensive rats subjected to permanent middle cerebral artery occlusion. Interleukin-1ß mRNA was quantified by Northern blot analysis and compared with rat macrophage RNA standard. To correct for gel loading, blots were also analyzed with cyclophilin cDNA, which encodes an abundant, conserved protein that was unchanged by the experimental conditions. Results: Interleukin-1ß mRNA produced in the ischemic zone was significantly increased from 6 hours to 120 hours, with a maximum of211±24% ofinterleukin-1ß reference standard, ie, 0.2 ng stimulated rat macrophage RNA, mRNA compared with the level in nonischemic cortices (4±2%) at 12 hours after ischemia (P<.OI; n=6). Interleukin-1ß mRNA at 12 hours after ischemia was markedly elevated in hypertensive rats over levels found in two normotensive rat strains. Neurological deficits were also apparent only in the hypertensive rats. Conclusions: Brain interleukin-1ß mRNA is elevated acutely after permanent focal ischemia and especially in hypertensive rats. These data suggest that this potent proinflammatory and procoagulant cytokine might have a role in brain damage following ischemia.
Feurestein
T Liu, P C McDonnell, P R Young, R F White, A L Siren, J M Hallenbeck, F C Barone and G Z
Interleukin-1 beta mRNA expression in ischemic rat cortex.
Print ISSN: 0039-2499. Online ISSN: 1524-4628
Copyright © 1993 American Heart Association, Inc. All rights reserved.
is published by the American Heart Association, 7272 Greenville Avenue, Dallas, TX 75231Stroke doi: 10.1161/01.STR.24.11.1746
1993;24:1746-1750Stroke.
http://stroke.ahajournals.org/content/24/11/1746
World Wide Web at:
The online version of this article, along with updated information and services, is located on the
http://stroke.ahajournals.org//subscriptions/ is online at: Stroke Information about subscribing to Subscriptions:
http://www.lww.com/reprints
Information about reprints can be found online at: Reprints:
document. Permissions and Rights Question and Answer process is available in the
Request Permissions in the middle column of the Web page under Services. Further information about this
Once the online version of the published article for which permission is being requested is located, click
can be obtained via RightsLink, a service of the Copyright Clearance Center, not the Editorial Office.Strokein Requests for permissions to reproduce figures, tables, or portions of articles originally publishedPermissions:
by guest on June 6, 2013http://stroke.ahajournals.org/Downloaded from
1746
Interleukin-1/3
mRNA
Expression
in
Ischemic
Rat
Cortex
T.
Liu,
MD;
P.C.
Mc
Donnell,
MS;
P.R.
Young,
PhD;
R.F.
White,
BA;
A.L.
Siren,
MD,
PhD;
J.M.
Hallenbeck,
MD;
F.C.
Barone,
PhD;
G.Z.
Feuerstein,
MD
Background
and
Purpose:
Interleukin-1lj
is
a
proinflammatory
cytokine
produced
by
blood-borne
and
resident
brain
inflammatory
cells.
The
present
study
was
conducted
to
determine
if
interleukin-13
mRNA
was
produced
in
the
brain
of
rats
subjected
to
permanent
focal
ischemia.
Methods:
Rat
interleukin-lj
cDNA,
synthesized
from
stimulated
rat
peritoneal
macrophage
RNA
by
reverse
transcription
and
polymerase
chain
reaction
and
cloned
in
plasmid
Bluescript
KS+,
was
used
to
evaluate
the expression
of
interleukin-1.3
mRNA
in
cerebral
cortex
from
spontaneously
hypertensive
rats
and
normotensive
rats
subjected
to
permanent
middle
cerebral
artery
occlusion.
Interleukin-1f3
mRNA
was
quantified
by
Northern
blot
analysis
and
compared
with
rat
macrophage
RNA
standard.
To
correct
for
gel
loading,
blots
were
also
analyzed
with
cyclophilin
cDNA,
which
encodes
an
abundant,
conserved
protein
that
was
unchanged
by
the
experimental
conditions.
Results:
Interleukin-10
mRNA
produced
in
the
ischemic
zone
was
significantly
increased
from
6 hours
to
120
hours,
with
a
maximum
of
211±24%
of
interleukin-lf
reference
standard,
ie,
0.2
ng
stimulated
rat
macrophage
RNA,
mRNA
compared
with
the
level
in
nonischemic
cortices
(4±2%)
at
12
hours
after
ischemia
(P<.01;
n=6).
Interleukin-1f3
mRNA
at
12
hours
after
ischemia
was
markedly
elevated
in
hypertensive
rats
over
levels
found
in
two
normotensive
rat
strains.
Neurological
deficits
were
also
apparent
only
in
the
hypertensive
rats.
Conclusions:
Brain
interleukin-1j3
mRNA
is
elevated
acutely
after
permanent
focal
ischemia
and
especially
in
hypertensive
rats.
These
data
suggest
that
this
potent
proinflammatory
and
procoagulant
cytokine
might
have
a
role
in
brain
damage
following
ischemia.
(Stroke.
1993;24:1746-1751.)
KEY
WoRDs
*
cerebral
ischemia
*
cytokines
*
neuronal
damage
*
rats
I
nterleukin-1,3
(IL-1,3)
is
a
cytokine
with
multiple
proinflammatory,
procoagulant,
and
cell
growth
modulatory
actions.'
The
presence
of
IL-l,3
in
the
central
nervous
system
is
believed
to
reflect
synthesis
by
diverse
cells
such
as
endothelium,
microglia,
astrocytes,
and
neurons.2
Interleukin-1,8
acts
via
specific
brain
IL-1
receptors
that
demonstrate
significant
spatial
distribution.3
As
in
peripheral
organs,
the
proinflammatory
and
prothrom-
botic
actions
of
IL-1f31
are
expected to
promote
acute
neuropathological
changes
at
excessive
levels
in
the
brain.
Brain
ischemia
has
been
shown
to
be
associated
with
an
acute
inflammatory
response,4-6
but
the
nature
of
the
inflammatory
mediators
involved
in
brain
isch-
emia
is
still
unknown.
While
numerous
reports
have
dealt
with
the
putative
roles
of
various
inflammatory
mediators
such
as
thromboxane
A2,
leukotrienes,
and
Received
August
14,
1992;
final
revision
received
April
26,
1993;
accepted
May
17,
1993.
From
the
Departments
of
Cardiovascular
Pharmacology
(T.L.,
R.F.W.,
F.C.B.,
G.Z.F.)
and
Molecular
Genetics
(P.C.Mc
D.,
P.R.Y.),
SmithKline
Beecham
Pharmaceuticals,
King
of
Prussia,
Pa;
and
the
Department
of
Neurology,
Uniformed
Services
Uni-
versity
of
the
Health
Sciences
(A.L.S.),
and
the
Stroke
Branch,
National
Institute
of
Neurological
Disorders
and
Stroke
(J.M.H.),
National
Institutes
of
Health,
Bethesda,
Md.
Correspondence
to
Giora
Z.
Feuerstein,
MD,
Department
of
Cardiovascular
Pharmacology,
SmithKline
Beecham
Pharmaceu-
ticals,
709
Swedeland
Rd,
PO
Box
1539,
King
of
Prussia,
PA
19406.
See
Editorial
Comment,
page
1750
platelet-activating
factor78
in
brain
injury,
few
attempts
have
been
made
to
explore
the
role
of
proinflammatory
cytokines
in
brain
injury.
Evidence
supporting
the
involve-
ment
of
cytokines
in
central
nervous
system
injury
in-
cludes
demonstrations
of
IL-1l3
mRNA
expression
in
mouse
brain
3
hours
after
endotoxin
administration,9
release
of
IL-1
in
mechanically
injured
brain
produced
by
intraparenchymal
implantation
of
a
microdialysis
probe,10
expression
of
IL-1,8
mRNA
in
rat
brain
in
response
to
direct
intraparenchymal
administration
of
the
neurotoxin
kainate,"1
or
combination
of
endotoxin
and
y-interferon.12
IL-1i8
mRNA
expression
was
also
shown
in
a
rat
model
of
transient
global
brain
ischemia
induced
by
permanent
bilateral
vertebral
artery
occlusion
followed
by
bilateral
carotid
occlusion
with
reperfusion.13
However,
cytokine
transcription
in
permanent
focal
ischemia
has
not
been
previously
evaluated.
The
purpose
of
the
present
study
was
to
determine
whether
the
initiation
phase
of
IL-1f3
production,
ie,
the
transcription
of
its
mRNA,
takes
place
after
permanent
middle
cerebral
artery
occlusion
(MCAO);
furthermore,
this
study
also
aimed
to
explore
the
differences,
if
any,
between
rats
carrying
risk
factors
for
stroke
(eg,
hypertension)
and
normotensive
rats.
Materials
and
Methods
To
provide
IL-1l8
mRNA
for
use
as
a
standard
and
to
create
the
necessary
IL-1,3
cDNA
probe
for
Northern
by guest on June 6, 2013http://stroke.ahajournals.org/Downloaded from
Liu
et
al
IL-13
mRNA
in
Focal
Ischemia
1747
blot
analysis,
rat
peritoneal
macrophages
were
collected
by
lavage
with
phosphate-buffered
saline
and
incubated
with
5
,g/mL
Escherichia
coli
lipopolysaccharide
(Sig-
ma
Chemical
Co,
St
Louis,
Mo)
at
37°C
for
4
hours.
Cells
were
lysed
in
4
mol/L
guanidinium
thiocyanate,
and
total
RNA
was
isolated
by
centrifugation
over
5.8
mol/L
(2sCl.'4
IL-1p
cDNA
was
synthesized
by
reverse
transcription
and
polymerase
chain
reactions
(RT-PCR;
Boehringer
Mannheim
Biochemica,
Indianapolis,
Ind)
using
the
5'
and
3'
synthetic
oligonucleotide
primers
5
'-CAGCGGCCGCCTTGTGCAAGTGTCTGAAG-
CAG-3'
and
3'-GAACAGCTCTTACCCGTCAGAG-
GTCCAGATCTCGCCGGCGAC-5',
respectively.
The
primers
were
engineered
to
contain
a
Noti
restriction
site
for
cloning
purposes.
The
fragment
obtained
by
RT-PCR
was
digested
with
Noti
and
ligated
into
Blue-
script
KS+
(Stratagene,
La
Jolla,
Calif).
The
identity
of
the
cDNA
insert
was
confirmed
by
sequencing
and
agrees
with
the
known
sequence.12
For
Northern
blot
hybridization,
the 0.9-kb
rat
IL-13
cDNA
insert
was
used
(isolated
by
Noti
digestion
of
Bluescript
K+
that
contained
the
IL-1,8
cDNA
fragment
and
purified
by
agarose
gel
electrophoresis).
Similarly,
a
human
0.9-kb
EcoRI
cyclophilin
cDNA
fragment
was
prepared
and
also
used
for
hybridization
according
to
Bergsma
et
al.15
Both
rat
IL-113
and
cyclophilin
cDNA
fragments
were
labeled
using
an
oligolabeling
kit
(Pharmacia
LKB
Biotechnology,
Piscataway,
NJ)
via
random
primer
ex-
tension
using
Klenow
DNA
Polymerase
I
and
purified
using
a
Nuctrap
push
column
(Stratagene).
Focal
ischemia
or
sham
surgery
was
carried
out
in
spontaneously
hypertensive
rats
(SHR;
Taconic
Farms,
Germantown,
NY)
and
in
two
normotensive
rat
strains
(F-344
or
Wistar-Kyoto;
Taconic
Farms)
(weight,
250
to
300
g).
Animals
were
anesthetized
with
sodium
pento-
barbital
(Steris
Laboratories,
Inc,
Phoenix,
Az;
60
mg/kg
IP)
and
prepared
for
surgery
as
previously
described.6
Briefly,
via
a
2-
to
3-mm
craniotomy
the
right
middle
cerebral
artery
was
isolated
on
the
hooked
tip
of
a
platinum-irridium
wire
(0.0045
-in
diameter;
Medwire,
Mount
Vernon,
NY)
and
then
simultaneously
occluded
and
cut
dorsal
to
the
lateral
olfactory
tract
(n=
18,
SHR;
n=4,
F-344;
n=4,
Wistar-Kyoto).
In
sham-operated
animals
(n=3;
SHR)
the
same
surgical
procedures
were
exercised,
but
the
artery
was
not
occluded,
and
animals
were
killed
at
12
hours.
In
addition,
nonoperated
animals
(n=3;
SHR)
were
also
studied.
Two
separate
neurological
examinations
were
per-
formed
on
the
rats
before
euthanasia
to
determine
the
severity
of
motor
deficits.
Contralateral
forelimb
def-
icits
were
measured
using
a
neurological
grade
as
previously
described,16
while
contralateral
hind-limb
deficits
were
measured
using
a
standard
hind-limb
placement
test.6
At
various
times
post-MCAO
rats
were
overdosed
with
pentobarbital,
and
forebrains
were
removed
and
dissected
as
described
previously.5'6
A
segment
of
the
ipsilateral
frontoparietal
cortex
was
sliced
from
the
hemisphere;
an
identical
segment
was
sliced
from
the
contralateral
cortical
hemisphere
(nonischemic
control).
The
segments
were
immedi-
ately
frozen
in
liquid
nitrogen
and
stored
at
-80°C.
For
Northern
blot
analysis
of
IL-1,3
mRNA
in
the
focal
ischemic
and
nonischemic
tissue,
total
RNA
from
peritoneal
macrophages
(ie,
0.2
,ug
isolated
as
described
above
and
used
in
the
Northern
blots
as
an
IL-1X3
mRNA
reference
standard
for
quantitation
purposes)
was
fractioned
through
a
1%
agarose
gel,
containing
6%
formaldehyde
in
MOPS
(3-[N-morpholino-]propane-
sulfonic
acid;
Sigma).
The
RNA
was
then
transferred
onto
a
nitrocellulose
membrane
(Keene,
NH;
BA83,
0.2
,um)
by
capillary
blotting
in
20x
saline-sodium
phos-
phate-ethylenediaminetetraacetic
acid
(EDTA)
(SSPE)
(lx
SSPE
is
150
mmol/L
NaCl,
2
mmol/L
NaH2P04,
and
1
mmol/L
EDTA).
RNA
was
immobi-
lized
on
the
filter,
prehybridized,
and
then
hybridized
to
the
32P-deoxycytidine
triphosphate-labeled
rat
IL-1p
cDNA
probes
at
42°C
for
14
to
18
hours
in
the
buffer
containing
50%
formamide,
6x
SSPE,
5x
Denhardt's
solution,
0.5%
sodium
dodecyl
sulfate
(SDS),
100
,g/mL
denatured
salmon
sperm
DNA,
and
10%
dex-
trane
sulfate
and
then
washed
to
a
final
stringency
of
0.5x
SSPE,
0.1%
SDS,
at
50°C
for
15
minutes.
The
membrane
was
exposed
to
Amersham
Hyperfilm-MP
for
2
days
with
intensifying
screen
and
developed
by
a
Kodak
M35A-OMAT
processor.
All
the
membranes
were
stripped
in
boiling
0.05
x
SSPE
and
rehybridized
to
32P-deoxycytidine
triphosphate-labeled
human
cyclo-
philin
cDNA
to
control
for
differences
in
RNA
loading.
Radioactivity
of
the
hybridized
blots
was
counted
by
a
Betascope
603
blot
analyzer
(Betagen
Corp,
Waltham,
Mass).
Cortical
tissue
IL-lp
mRNA
was
quantitated
for
between-blot
comparisons
as
the
percent
relative
radio-
activity
of
the
reference
standard
(ie,
macrophage
RNA
loaded
on
the
same
gel
normalized
to
the
actual
amount
of
RNA
loaded
that
was
determined
from
the
cyclo-
philin
mRNA
counts
from
the
same
cortical
samples).
For
within-blot
comparisons
of
ischemic
cortex
IL-1f3,
percent
relative
radioactivity
was
normalized
directly
in
relation
to
the
loaded
RNA
determined
from
the
cyclo-
philin
mRNA
in
each
sample.
All
data
are
expressed
as
mean±SEM.
Statistical
analysis
of
the
data
was
performed
using
analysis
of
variance
and
Tukey's
multiple
comparison
tests.
Signif-
icant
differences
were
accepted
at
P<.05.
Results
Neurological
deficits
after
focal
ischemia
occurred
in
SHR
(Fig
1)
but
not
in
normotensive
animals.
Both
forelimb
and
hind-limb
dysfunction
in
SHR
were
ob-
served.
Partial
spontaneous
recovery
was
evidenced
at
5
days
for
forelimb
scores
and
at
1
and
5
days
for
hind-limb
scores.
Fig
2
illustrates
the
results
of
a
representative
North-
ern
blot
of
rat
cortical
RNA
12
hours
after
MCAO.
IL-1,8
mRNA
is
significantly
more
abundant
in
ipsilat-
eral
(top
IL-1,3
lanes
1
through
6)
ischemic
cortices
than
in
the
contralateral
cortices
(bottom
IL-1p
lanes
1
through
6)
of
the
same
animals.
RNA
from
macro-
phages
was
used
as
the
reference
standard
for
quanti-
tation
purposes
(Fig
2);
correcting
for
total
loaded
RNA
was
done
by
measurement
of
the
cyclophilin
mRNA
(cyclophilin
lanes
1
through
6)
from
the
ischemic
(top)
and
nonischemic
(bottom)
regions.
Fig
3
depicts
the
quantitated
IL-lp
mRNA
levels
in
the
cortices
taken
at
various
time
points
after
MCAO.
IL-1i8
mRNA
was
not
detectable
in
cortical
samples
from
nonoperated
rats
(0±0%,
n=3).
Sham
surgery
produced
no
significant
effect.
In
the
contralateral
cortical
tissues
(50
lxg)
and
total
RNA
obtained
from
by guest on June 6, 2013http://stroke.ahajournals.org/Downloaded from
1748
Stroke
Vol
24,
No
11
November
1993
2.5
-
0
C1
n
jo
0
IL
2.0
-
1.5
-
1.0
-
0.5
1.2
-
0
'm
C
I
E
V5
.r
1.0
0.8
-
0.6
-
0.4
-
0.2
-
*
*
*
Sham
6
hr
12
hr
24
hr
5
day
T
T
0.0-
Sham
6
hr
12
hr
24
hr
5
day
FIG
1.
Bar
graphs
illustrate
graded
forelimb
(top)
and
hind-limb
(bottom)
deficits
in
sham
(n-3)
and
permanent
middle
cerebral
artery
occlusion
(n=3
to
6)
in
spontane-
ously
hypertensive
rats.
*P<.05
compared
with
sham.
Neurological
deficits
at
1
hour
after
middle
cerebral
artery
occlusion
could
not
be
determined
because
of
anesthe-
sia.
Vertical
bar
lines
represent
SEM.
nonischemic
cortex
after
MCAO,
a
weak
trend
for
expression
of
IL-i83
was
detected
that
did
not
reach
statistical
significance.
Permanent
MCAO
dramatically
increased
the
IL-1f8
mRNA
level
in
the
ipsilateral
ischemic
cortex
compared
with
the
contralateral
control
cortex
from
6
hours
after
MCAO
and
up
to
5
days
later.
Peak
IL-1if
mRNA
expression
was
noticed
12
hours
after
MCAO
(P<.01),
but
significantly
elevated
levels
were
still
monitored
5
days
later
(P<.05).
To
examine
whether
the
elevated
levels
of
IL-1i8
found
in
the
ischemic
cortices
are
specific
to
SHR
only,
we
have
monitored
IL-1i6
mRNA
in
three
groups
(n=4
per
group)
of
rats
studied
simultaneously
(ie,
killed
at
12
hours
after
MCAO
and
quantitated
together).
The
two
normotensive
strains,
WKY
and
F-344
rats,
responded
with
much
less
IL-Ip
mRNA
expression
compared
with
the
SHR
group
(Fig
4).
However,
the
levels
of
IL-1,i
mRNA
in
both
normotensive
rat
strains
were
signifi-
cantly
elevated
over
background
levels.
Discussion
This
article
demonstrates
that
IL-1if
mRNA
in
rat
brain
is
significantly
increased
in
ischemic
cerebral
lschemic
Cortex
1
2
3
4
5 6
7
IL-1,
Ps
Cyclophilin
w
*
Nonischemic
Cortex
1
2 3
4
5 6
7
IL-1l
S
Cyclophilin
50i
@
0
M
*
FIG
2.
Northern
blot
of
interieukin-1,/
(IL-1i8)
from
cere-
bral
cortices
after
12-hour
permanent
middle
cerebral
artery
occlusion
(n=6).
Lanes
1
through
6,
50
gg
of
total
RNA
from
ipsilateral
(ischemic)
or
contralateral
(nonisch-
emic)
cortex
was
separated
on
1
%
agarose
gel
contain-
ing
6%
formaldehyde,
blotted
into
nitrocellulose
mem-
brane,
and
then
hybridized
with
rat
cDNA
IL-1p8
probe.
Lane
7,
0.2
gg
of
total
RNA
was
isolated
from
rat
macrophages
stimulated
with
5
gg/mL
lipopolysaccha-
ride
loaded
on
the
same
gel.
After
stripping,
the
blot
was
hybridized
with
a
human
cyclophilin
cONA
probe
and
used
in
quantitation
of
the
data
to
control
for
amount
of
loaded
RNA.
tissue.
The
time
course
of
IL-1:
mRNA
expression
demonstrated
a
trend
toward
increased
IL-1/3
mRNA
expression
as
early
as
1
hour
after
permanent
MCAO,
a
definite
increase
at
6
hours,
and
persistently
elevated
IL-1,3
mRNA
levels
up
to
5
days
after
ischemia.
The
early
IL-1p
mRNA
expression,
ie,
1
to
6
hours,
pre-
cedes
the
time
of
leukocyte
infiltration
(measured
by
mycloperoxidase
accumulation)
from
blood
vessels
out-
side
the
ischemic
zone,
which
only
begins
to
occur
after
12
hours
and
then
dramatically
increases
for
a
period
of
5
days
after
infarction
in
this
model.'7
These
data
are
further
supported
by
a
detailed
histological
study
show-
ing
significant
elevation
of
neutrophils
in
ischemic
cor-
tex
at
48
hours
after
ischemia.'8
Our
data
also
suggest
that
IL-1,B
mRNA
expression
in
focal
ischemic
brain
tissue
is
not
specific
to
the
SHR
strain
since
significant
IL-1:3
mRNA
expression
has
been
clearly
demonstrated
in
the
two
normotensive
rat
strains.
The
lesser
IL-1J3
mRNA
expression
in
the
two
normotensive
rat
strains
may
represent
the
lesser
vulnerability
of
brain
cortex
of
normotensive
rats
to
permanent
MCAO,6
as
also
re-
flected
in
the
present
study
by
lack
of
neurological
deficits
in
the
normotensive
rats
and
previous
reports
on
infarct
volume.6
The
data
presented
in
this
study
are
in
accord
with
recent
reports
demonstrating
IL-1,3
mRNA
expression
in
several
different
brain
injury
models.9-'3
Thus,
Mi-
nami
et
al13
have
shown
acute
biphasic
IL-18
mRNA
expression
in
various
brain
regions
after
transient
global
ischemia.
However,
several
important
differences
be-
tween
the
two
studies
must
be
pointed
out.
First,
the
present
data
represent
transcription
of
IL-1f3
in
perma-
nent
focal
ischemia.
Second,
the
previous
study'3
has
0.0
by guest on June 6, 2013http://stroke.ahajournals.org/Downloaded from
Liu
et
al
IL-1f
mRNA
in
Focal
Ischemia
1749
300 -
>.h
0
0
cc
._
0
._
cc
z
Cc
250
-
200
-
150
-
100
-
50
-
*
ischemic
cortex
E
nonischemic
cortex
**
I
**
.
*
0
1
=
Z
WR
E
:
Sham
lhr
MCAO
6hr
MCAO
12hr
MCAO
24hr
MCAO
5day
MCAO
FIG
3.
Bar
graph
illustrates
quantitative
analysis
of
the
time
course
for
interleukin-1,p
(IL-1
p)
mRNA
production
in
cerebral
cortices
ipsilateral
or
contralateral
to
permanent
middle
cerebral
artery
occlusion
(MCAO)
(n=3
to
6
per
time
point)
or
sham
surgery
(n=3
killed
at
12
hours).
The
}-emissions
from
bands
on
hybridized
Northern
blots
were
counted
using
a
Betascope
counter.
Data
(mean+±SE)
from
different
blots
are
presented
as
percentage
of
loaded
rat
macrophage
IL-1
p
mRNA
(reference
standard)
normalized
to
the
actual
amount
of
loaded
RNA
(see
"Materials
and
Methods").
*P<.05
compared
with
contralateral
cortex;
**P<.01
compared
with
contralateral
cortex
and
compared
with
ipsilateral
cortex
of
5-day
MCAO.
No
detectable
IL-1,8
mRNA
was
observed
in
either
cortex
of
nonoperated
rats
(n=3;
data
not
shown).
Vertical
bars
represent
SEM.
not
used
a
reference
message
(eg,
cyclophilin),
nor
did
it
standardize
its
hybridization
probe
against
a
single
IL-lp
mRNA-containing
sample
such
as
rat
macrophage
IL-113
mRNA.
Third,
and
most
importantly,
the
present
study
provides
quantitated
data,
whereas
only
qualitative
(visu-
al)
representative
gels
were
provided
in
the
global
isch-
emia
study.13
These
differences
may
underlie
the
lack
of
significant
IL-1,8
mRNA
in
the
brain
between
1
and
7
days
after
global
brain
ischemia,
whereas
significant
IL-1i8
mRNA
was
clearly
demonstrated
in
our
study
at
5
days
1400
-
C)
0
a)
1200-
1i000
800
600
I
400
_
_
~~~**
200-
SHR
WKY
Fisher
FIG
4.
Bar
graph
illustrates
quantitative
analysis
of
isch-
emic
cortex
interleukin-1,p
mRNA
production
in
groups
(n=4)
of
hypertensive
(spontaneously
hypertensive
rats
[SHR])
and
normotensive
(Wistar-Kyoto
[WKY]
and
Fisher)
rats
12
hours
after
middle
cerebral
artery
occlu-
sion.
Data
from
one
blot
are
presented
as
relative
change
normalized
to
the
actual
amount
of
loaded
RNA
(see
"Materials
and
Methods").
*P<.05
compared
with
SHR;
**P<.05
compared
with
WKY
rats.
after
MCAO.
The
lack
of
standardization
and
quantifica-
tion
of
IL-1X
mRNA
in
the
global
ischemia
report
also
makes
it
difficult
to
correlate
the
magnitude
of
the
IL-1p
mRNA
expression
in
the
focal
ischemic
brain
cortex
described
here
to
the
changes
described
previously
in
the
global
ischemia
model.13
The
significance
of
the
IL-18
mRNA
expression
de-
scribed
in
the
present
study
must
be
interpreted
with
caution
because
no
evidence
has
been
provided
to
indicate
that
translation
of
the
message
into
the
functional
cyto-
kine
has
occurred
either
at
the
time
of
the
IL-1,8
mRNA
expression
or
thereafter.
However,
the
de
novo
expression
of
IL-1p
mRNA
in
the
nonperfused
ischemic
cortex
supports
the
possibility
that
endogenous
IL-lp
might
be
produced
in
ischemic
brain.
This
possibility
draws
further
credence
from
reports
demonstrating
the
capacity
of
the
brain
to
synthesize
cytokines,
including
IL-1,
in
several
other
brain
injury
models.
For
example,
mechanical
inju-
ry'0
resulted
in
significant
IL-1
production
24
hours
after
injury,
which
compares
favorably
with
the
earlier
IL-1,3
mRNA
shown
in
our
study.
Also
in
this
latter
study,
IL-1
production
was
significantly
lower
at
7
days
after
injury,
again
in
accord
with
the
significant
diminution
of
IL-1ij
mRNA
observed
in
our
study.
In
any
case,
it
will
be
necessary
to
directly
monitor
IL-1p
production
in
focal
brain
ischemia
to
precisely
relate
the
transcriptional
event
to
the
cytokine
production.
The
cellular
elements
expressing
IL-1,i
mRNA
in
the
ischemic
brain
have
not
been
elucidated
in
the
present
study.
However,
IL-1
has
been
shown
to
be
produced
by
microglial
cells
in
vitro19
and
in
vivo2021
and
by
astro-
cytes
in
vitro.22,23
Furthermore,
IL-1
has
been
detected
in
the
cerebrospinal
fluid
of
rats.24
It
is
also
noteworthy
that
a
peripheral
source
for
IL-1
might
gain
access
into
the
brain
since
this
cytokine
was
shown
to
be
trans-
ported
from
the
blood
into
the
brain25;
therefore,
even
i..-
-
by guest on June 6, 2013http://stroke.ahajournals.org/Downloaded from
1750
Stroke
Vol
24,
No
11
November
1993
if
IL-118
is
produced
in
the
brain,
this
may
not
preclude
a
role
for
peripheral
IL-1
in
brain
injury.
If
one
accepts
the
possibility
that
the
increase
in
IL-1p
mRNA
shortly
after
ischemic
stroke
indeed
translates
into
a
robust
production
of
this
potent
cyto-
kine,
one
must
also
anticipate
its
consequences;
these
include
activation
of
microglia
and
astrocytes26
and
transformation
of
the
endothelium
into
a
proinflamma-
tory
and
prothrombotic27
state
that
may
propagate
thrombosis
and
the
extension
of
the
ischemic
zone.
Finally,
it
is
important
to
point
out
that
IL-1p
mRNA
is
only
one
of
many
genes
that
are
overexpressed
in
the
brain
after
injury;
most
notably,
acute immediate
genes
such
as
C-foS,28
heat-shock
protein,2829
and
the
tumor
suppressor
gene
product
p5330
have
been
shown
to
be
overexpressed
in
the
brain
in
response
to
ischemia.
References
1.
Dinarello
CA.
Biology
of
interleukin-1.
FASEB
J.
1988;2:108-115.
2.
Rothwell
NJ.
Functions
and
mechanism
of
interleukin-1
in
the
brain.
Trends
Pharmacol
Sci.
1991;12:430-435.
3.
Takao
T,
Tracey
DE,
Mitchel
WM,
DeSouza
EB.
Interleukin-1
receptors
in
mouse
brain:
characterization
and
neuronal
local-
ization.
Endocrinology.
1990;127:3070-3078.
4.
Barone
FC,
Hillegass
LM,
Price
WJ,
White
RF,
Lee
EV,
Feu-
erstein
GZ,
Sarau
HM,
Clark
RK,
Griswold
DE.
Polymorpho-
nuclear
leukocyte
infiltration
into
cerebral
focal
ischemic
tissue:
myeloperoxidase
activity
assay
and
histologic
verification.
J
Neurosci
Res.
1991;29:336-345.
5.
Barone
FC,
Schmidt
DB,
Hillegass
LM,
Price
WJ,
White
RF,
Feuerstein
GZ,
Clark
RK,
Lee
EV,
Griswold
DE,
Sarau
HM.
Reperfusion
increases
neutrophils
and
LTB4
receptor
binding
in
rat
focal
ischemia.
Stroke.
1992;23:1337-1348.
6.
Barone
FC,
Price
WJ,
White
RF,
Willette
RN,
Feuerstein
GZ.
Genetic
hypertension
and
increased
susceptibility
to
cerebral
ischemia.
Neurosci
Biobehav
Rev.
1992;16:219-233.
7.
Krieglstein
J.
Pharmacology
of
Cerebral
Ischemia.
New
York,
NY:
Elsevier
Science
Publishing
Co,
Inc;
1986.
8.
Jane
JA,
Anderson
DK,
Torner
JC,
Young
W.
Central
nervous
system
trauma
status
report.
J
Neurotrauma.
1992;9:S1-S407.
9.
Ban
E,
Haour
F,
Lenstra
R.
Brain
interleukin-1
expression
induced
by
peripheral
lipopolysaccharide
administration.
Cytokine.
1992;4:48-54.
10.
Woodroofe
MN,
Sarna
GS,
Wadhwa
M,
Hayes
GM,
Loughlin
AJ,
Tinker
A,
Cuzner
ML.
Detection
of
interleukin-1
and
interleukin-6
in
adult
rat
brain,
following
mechanical
injury,
by
in
vivo
macrodialysis:
evidence
of
a
role
for
microglia
in
cytokine
production.
JNeuroimmunol.
1991;33:227-236.
11.
Minami
M,
Kuraishi
Y,
Satoh
M.
Effects
of
kainic
acid
on
mes-
senger
RNA
levels
of
IL-1,3,
IL-6,
TNFa
and
LIF
in
the
rat
brain.
Biochem
Biophys
Res
Commun.
1991;76:593-598.
12.
Higgins
GA,
Olschowka
JA.
Induction
of
interleukin-1,3
on
RNA
in
adult
rat
brain.
Mol
Brain
Res.
1991;9:143-148.
13.
Minami
M,
Kuraishi
Y,
Yabuuchi
K,
Yamazaki
A,
Satoh
M.
Induction
of
interleukin-lp
mRNA
in
rat
brain
after
transient
forebrain
ischemia.
J
Neurochem.
1991;58:390-392.
14.
Sambrook
J,
Fritsch
EF,
Maniatis
T.
Molecular
Cloning:
A
Labo-
ratory
Manual.
Cold
Spring
Harbor,
NY:
Cold
Spring
Harbor
Lab-
oratory
Press;
1989.
15.
Bergsma
DJ,
Eder
C,
Gross
M,
Kersten
H,
Sylvester
D,
Applebaum
E,
Cusimaro
D,
Livi
GP,
McLaughlin
MM,
Kasyan
K,
Porter
TG,
Silverman
C,
Dunningham
D,
Hand
A,
Prichett
WP,
Bossard
MJ,
Brant
M,
Levy
MA.
The
cyclophilin
multigene
family
of
peptidyl-prolyl-isomerases:
characterization
of
three
separate
isoforms.
J
Biol
Chem.
1991;266:23204-23214.
16.
Bederson
JB,
Pitts
LH,
Nishimura
MC,
Davis
RL,
Bartkowski
H.
Rat
middle
cerebral
artery
occlusion:
evaluation
of
the
model
and
devel-
opment
of
a
neurological
examination.
Stroke.
1986;17:472-476.
17.
Clark
RK,
Lee
EV,
Fish
CJ,
White
RF,
Price
WJ,
Jonak
ZL,
Feuerstein
GZ,
Barone
FC.
Development
of
tissue
damage,
inflammation
and
resolution
following
stroke:
an
immunohisto-
chemical
and
quantitative
planimetric
study.
Brain
Res
Bull.
1993;
31:565-572.
18.
Dereski
MO,
Chopp
M,
Knight
RH,
Chen
H,
Garcia
JH.
Focal
cerebral
ischemia
in
the
rat:
temporal
profile
of
neutrophil
responses.
Neurosci
Res
Commun.
1992;11:179-186.
19.
Hetier
E,
Dene
HE,
Bousseau
A,
Rouget
P,
Mallat
M,
Prochiantz
A.
Brain
macrophages
synthesize
interleukin-1
and
interleukin-1
mRNA
in
vitro.
J
Neurosci
Res.
1988;21:391-397.
20.
Fontana
A,
Weber
E,
Dayer
JM.
Synthesis
of
interleukin-1/
endogenous
pyrogen
in
the
brain
of
endotoxin
treated
mice:
a
step
in
fever
induction.
J
ImmunoL
1984;133:1696-1698.
21.
Guilian
D,
Baker
TJ,
Shih
L,
Lachman
LB.
Interleukin-1
of
the
central
nervous
system
is
produced
by
ameboid
microglia.
J
Exp
Med.
1986;164:595-603.
22.
Fontana
A,
Kristensen
F,
Dules
R,
Gemsa
D,
Weber
E.
Pro-
duction
of
prostaglandin
E
and
interleukin-1
like
factor
by
cultured
astrocytes
and
C-
6
glioma
cells.
J
Immunol.
1982;129:2413
-2419.
23.
Lieberman
AP,
Pitha
PM,
Shin
HS,
Shin
ML.
Production
of
tumor
necrosis
factor
and
other
cytokines
by
astrocytes
stimulated
with
lipopolysaccharide
or
a
neurotrophic
virus.
Proc
NatlAcad
Sci
USA.
1989;86:56348-56352.
24.
Coceani
F,
Lees
J,
Pinarello
CA.
Occurrence
of
interleukin-1
in
cerebrospinal
fluid
of
conscious
rat.
Brain
Res.
1988;446:245-250.
25.
Banks
WA,
Ortiz
L,
Kastin
AJ.
Human
interleukin
(IL)
l
a,
murine
IL-la
and
murine
IL-1,8
are
transported
from
blood
to
brain
in
the
mouse
by
a
shared
saturable
mechanism.
J
Pharmacol
Exp
Ther.
1991;259:988-996.
26.
Benveniste
EN.
Cytokines:
influence
on
glial
cell
gene
expression
and
function.
Neuroimmunoendocrinol.
1992;52:106-153.
27.
Montovani
A,
Bussolino
F,
Dejani
E.
Cytokine
regulation
of
endo-
thelial
cell
function.
FASEB
J.
1992;6:2591-2599.
28.
Nowak
TS
Jr,
Ikeda
J,
Nakajima
T.
72-kDa
Heat
shock
protein
and
c-fos
gene
expression
after
transient
ischemia.
Stroke.
1990;
21(suppl
3):I1I-107-III-111.
29.
Li
Y,
Chopp
M,
Garcia
JH,
Yoshida
Y,
Zhang
LG,
Levine
SR.
Distribution
of
the
72-kd
heat-shock
protein
as
a
function
of
transient
focal
cerebral
ischemia
in
rats.
Stroke.
1992;23:1292-1298.
30.
Chopp
M,
Zheng
YL,
Freytag
SO.
P
53
expression
in
brain
after
middle
cerebral
artery
occlusion
in
the
rat.
Biochem
Biophys
Res
Commun.
1992;182:1201-1207.
Editorial
Comment
Recent
advances
and
research
interest
in
molecular
biology
have
prompted
stroke
researchers
and
neurosci-
entists
to
study
gene
induction
and
expression
in
ischemic
brain.
A
good
example
of
this
interest
is
the
induction
and
expression
of
the
70-kD
stress
(heat-shock)
gene
at
both
transcriptional
and
translational
levels
in
neurons,
glia,
and
endothelial
cells
that
have
been
intensively
investi-
gated
in
various
models
of
permanent
and
temporary
cerebral
ischemia.1-6
Along
a
similar
line,
using
a
rat
interleukin
(IL)-1l3
cDNA
synthesized
from
stimulated
rat
peritoneal
macrophage
RNA,
Liu
et
al
have
now
reported
the
induction
and
expression
of
IL-1X3
mRNA
in
the
ischemic
zone
in
both
hypertensive
and
normotensive
rats
after
permanent
focal
ischemia.
Furthermore,
the
brain
level
of
IL-1X3
mRNA
is
increased
markedly
in
hyperten-
sive
rats
over
the
levels
found
in
two
different
strains
of
normotensive
rats.
A
couple
of
important
implications
resulting
from
this
interesting
study
might
need
elabora-
tion.
First,
the
elevated
levels
of
IL-1,3
mRNA
in
the
brain
cortex
in
hypertensive
rats
correlate
with
the
severity
of
the
neurological
deficits
after
permanent
focal
ischemia,
which
suggests
that
the
elevated
levels
of
cytokine
mRNA
by guest on June 6, 2013http://stroke.ahajournals.org/Downloaded from
... TNF-α share some of similarities to IL-1 and IL-6 in the development of inflammatory responses. In MCAO rat model, TNF-α increases significantly, proportionally to IL-1 and IL-6 levels, within hours post-IRI ( Liu et al., 1993). Increased levels of TNF-α in CSF and serum was reported to be neurotoxic, increasing infarct volume while decrease in its levels was neuroprotective ( Zaremba et al., 2001;Shi et al., 2018;Xu et al., 2018). ...
Article
Full-text available
Acquired brain ischemia-and reperfusion-injury (IRI), including both Ischemic stroke (IS) and Traumatic Brain injury (TBI), is one of the most common causes of disability and death in adults and represents a major burden in both western and developing countries worldwide. China’s clinical neurological therapeutic experience in the use of traditional Chinese medicines (TCMs), including TCM-derived active compounds, Chinese herbs, TCM formulations and decoction, in brain IRI diseases indicated a trend of significant improvement in patients’ neurological deficits, calling for blind, placebo-controlled and randomized clinical trials with careful meta-analysis evaluation. There are many TCMs in use for brain IRI therapy in China with significant therapeutic effects in preclinical studies using different brain IRI-animal. The basic hypothesis in this field claims that in order to avoid the toxicity and side effects of the complex TCM formulas, individual isolated and identified compounds that exhibited neuroprotective properties could be used as lead compounds for the development of novel drugs. China’s efforts in promoting TCMs have contributed to an explosive growth of the preclinical research dedicated to the isolation and identification of TCM-derived neuroprotective lead compounds. Tanshinone, is a typical example of TCM-derived lead compounds conferring neuroprotection toward IRI in animals with brain middle cerebral artery occlusion (MCAO) or TBI models. Recent reports show the significance of the inflammatory response accompanying brain IRI. This response appears to contribute to both primary and secondary ischemic pathology, and therefore anti-inflammatory strategies have become popular by targeting pro-inflammatory and anti-inflammatory cytokines, other inflammatory mediators, reactive oxygen species, nitric oxide, and several transcriptional factors. Here, we review recent selected studies and discuss further considerations for critical reevaluation of the neuroprotection hypothesis of TCMs in IRI therapy. Moreover, we will emphasize several TCM’s mechanisms of action and attempt to address the most promising compounds and the obstacles to be overcome before they will enter the clinic for IRI therapy. We hope that this review will further help in investigations of neuroprotective effects of novel molecular entities isolated from Chinese herbal medicines and will stimulate performance of clinical trials of Chinese herbal medicine-derived drugs in IRI patients.
... Inflammation is one of the major consequences of a stroke [33]. In the ischemic cortex, levels of pro-inflammatory mediators, including cytokines and adhesion molecules, increase about 1h after stroke, and return to basal values after 5 days [34, 35]. Thus, a milder inflammation could explain the smaller effect of our distant lesion. ...
Article
Full-text available
It was previously shown that a small lesion in the primary somatosensory cortex (S1) prevented both cortical plasticity and sensory learning in the adult mouse visual system: While 3-month-old control mice continued to show ocular dominance (OD) plasticity in their primary visual cortex (V1) after monocular deprivation (MD), age-matched mice with a small photothrombotically induced (PT) stroke lesion in S1, positioned at least 1 mm anterior to the anterior border of V1, no longer expressed OD-plasticity. In addition, in the S1-lesioned mice, neither the experience-dependent increase of the spatial frequency threshold ("visual acuity") nor of the contrast threshold ("contrast sensitivity") of the optomotor reflex through the open eye was present. To assess whether these plasticity impairments can also occur if a lesion is placed more distant from V1, we tested the effect of a PT-lesion in the secondary motor cortex (M2). We observed that mice with a small M2-lesion restricted to the superficial cortical layers no longer expressed an OD-shift towards the open eye after 7 days of MD in V1 of the lesioned hemisphere. Consistent with previous findings about the consequences of an S1-lesion, OD-plasticity in V1 of the nonlesioned hemisphere of the M2-lesioned mice was still present. In addition, the experience-dependent improvements of both visual acuity and contrast sensitivity of the open eye were severely reduced. In contrast, sham-lesioned mice displayed both an OD-shift and improvements of visual capabilities of their open eye. To summarize, our data indicate that even a very small lesion restricted to the superficial cortical layers and more than 3mm anterior to the anterior border of V1 compromised V1-plasticity and impaired learning-induced visual improvements in adult mice. Thus both plasticity phenomena cannot only depend on modality-specific and local nerve cell networks but are clearly influenced by long-range interactions even from distant brain regions.
... These processes all result in neuronal cell death and enhance the damage to the ischemic brain. Analysis of the temporal profile of mRNA expression of cytokines in ischemic rats, have revealed that the up-regulation of TNF-αmRNA is proportional to IL-1 and IL-6 up- regulation6465. Initial increases are seen 1-3 h after ischemia onset [66] , and have a twophase pattern of expression with a second peak at 24-36 h6768. In particular, the These three proinflammatory cytokines are able to affect the volume of the ischemic induced tissue damage in rodent experimental stroke [76]. ...
Article
Full-text available
Stroke is an major public health issue and third leading cause of death after heart disease. affecting 15 million people worldwide each year. Although extensive research on stroke has been carried out during past decade the current therapeutic strategies have been largely unsuccessful due to number of reasons. Probably the research and pharmacological management have focused on very early events in brain ischemia. Two important pathophysiological mechanisms involved during ischemic stroke are oxidative stress and inflammation. Brain tissue is not well equipped with antioxidant defenses, so reactive oxygen species and other free radicals/oxidants, released by inflammatory cells, threaten tissue viability in the vicinity of the ischemic core. Recent studies have shown that brain ischemia and trauma elicit strong inflammatory reactions driven by both external and brain cells. Clinical observations suggest that patients with stroke have higher plasma levels of inflammatory cytokines or soluble adhesion molecules and anti-inflammatory therapy is effective at reducing stroke incidence in not only animal models, but in humans as well. This suggests that inflammation might directly affect the onset of stroke. The recognition of inflammation as a fundamental response to brain ischemia provides novel opportunities for new anti-inflammatory therapies. Currently, little is known about endogenous counter regulatory immune mechanisms. Statins have been shown to decrease the stroke incidence via anti-inflammatory effects that are both dependent and independent of their cholesterollowering effects. Here in this review we will discuss the molecular aspects of oxidative stress and inflammation in ischemic stroke. We will also present the latest findings about the cellular and humoral aspects of immune and inflammatory reactions in the brain. This article may have an impact on the potential therapeutic strategies that target neuro-inflammation and the innate immune system.
... Cytokines are upregulated in the brain in response of a variety of stimulus including ischemia, being IL-1, interleukin- 6 (IL-6), TNF-a, interleukin-10 (IL-10) and TGF-b, the most studied cytokines related to inflammation in stroke [39] [40] [41] [42]. Analysis of the temporal profile of mRNA expression of cytokines in ischemic rats, have revealed that the up-regulation of TNF-αmRNA is proportional to IL-1 and IL-6 up-regulation [64] [65]. Initial increases are seen 1–3 h after ischemia onset [66], and have a two-phase pattern of expression with a second peak at 24–36 h [67] [68]. ...
Article
Full-text available
Stroke is an important public health issue due to high rates of disability, morbidity/mortality and is now the third leading cause of death after heart disease and cancer affecting 15 million people worldwide each year. In spite of extensive research in the field of stroke during past decade the current therapeutic strategies have been largely unsuccessful. One possible explanation is that research and pharmacological management have focused on very early events in brain ischemia. Two important pathophysiological mechanisms involved during ischemic stroke are oxidative stress and inflammation. Brain tissue is not well equipped with antioxidant defenses, so reactive oxygen species and other free radicals/oxidants, released by inflammatory cells, threaten tissue viability in the vicinity of the ischemic core. Recent studies have shown that brain ischemia and trauma elicit strong inflammatory reactions driven by both external and brain cells. Clinical observations suggest that patients with stroke have higher plasma levels of inflammatory cytokines or soluble adhesion molecules and anti-inflammatory therapy is effective at reducing stroke incidence in not only animal models, but in humans as well. This suggests that inflammation might directly affect the onset of stroke. The recognition of inflammation as a fundamental response to brain ischemia provides novel opportunities for new anti-inflammatory therapies. Currently, little is known about endogenous counter regulatory immune mechanisms. Statins have been shown to decrease the stroke incidence via anti-inflammatory effects that are both dependent and independent of their cholesterol-lowering effects. Here in this review we will discuss the molecular aspects of oxidative stress and inflammation in ischemic stroke. We will also present the latest findings about the cellular and humoral aspects of immune and inflammatory reactions in the brain. This will increase our understanding regarding neuro-injuries and role immune reactions play in the brain milieu. This all may have an impact on the potential therapeutic strategies that target neuro-inflammation and the innate immune system.
Thesis
Quantitative human mRNA data are derived from post-mortem or biopsied tissue. Confounding factors, RNA degradation, poor replication and a large variance are often cited, however, as objections to the data's reliability. At issue is whether post-mortem mRNA represents an ordered system and to what degree non-specific factors contribute to the measurements. I developed statistical methods and validated them by measuring 25 mRNA transcripts in an animal model of ischaemia. In the process I discovered novel increases for 3 genes in rats with ischaemic damage: leukaemia inhibitory factor, nestin and galanin mRNA. Additionally, I discovered that reference genes known as "housekeepers"' do not always act as steady-state controls and that the precise value of a test gene response varies according to the baseline choice of reference gene. Once optimised, I applied the analytical methods to human post-mortem brains. I used TaqMan™ real-time RT-PCR to measure 13 mRNAs in 513 cortical samples taken from 90 Alzheimer's disease and 81 control brains. Despite a high variance and confounding factors such as brain pH, I found strong geometric relations between the mRNA transcripts up to and beyond 100 hours autopsy delay. Where a postmortem brain had a high/low level of one mRNA, the same brain invariably had a high/low level of other mRNAs; correlated order is present and provides a means of isolating any mRNA change due specifically to disease. I measured mRNA levels of β-Secretase (BACE), GSK 3 and the isoforms of APP/APRP in the AD and control brains. After adjustment for age of death, brain pH, and gender, there was no change in the mRNA levels for either BACE or GSK 3α mRNA (p = 0.354 and p = 0.054 respectively). There was a change, however, in the ratio of KPI+ to KPI- mRNA isoforms of APP/APRP. Three separate probes, designed only to recognise KPI+ mRNA, each gave increases of between 28 and 50% in AD brains relative to controls (p = 0.002). There was no change in the mRNA levels of KPI-(APP 695) (p = 0.898). Therefore, whilst I KPI- mRNA levels remained level between AD and control brains, the KPI+ species were seen to increase specifically in the AD brains.
Chapter
Our results on the role of nitric oxide (NO) in cellular mechanisms of ischemic brain damage are as follows. (1) Repeated i.p. administration of N G-nitro-L-argininc (L-NNA) mitigated rat brain edema or infarction following permanent middle cerebral artery (MCA) occlusion at a dosage of 0.01–1 mg/kg. (2) In brain microvessels obtained from the affected hemisphere, Ca2+-dependent constitutive nitric acid synthase (NOS) (E-c-NOS) was activated consistently at an early stage. On the other hand, Ca2+-independent inducible NOS (i-NOS) was activated at 4h and 24h after MCA occlusion. (3) In the cerebral cortex, only Ca2+-dependent constitutive NOS (N-c-NOS) was activated during the first 4h of MCA occlusion in rats. (4) A2-h reperfusion following a 2-h MCA occlusion caused a significant increase in E-c-NOS and N-c-NOS activity without any apparent alterations in i-NOS activity. (5) NO concentration in the affected hemisphere was remarkably increased at 15–30 min and subsequently at 3–4 h after MCA occlusion. (6) Restoration of blood flow for 2h after a 2-h MCA occlusion resulted in an enhanced NO concentration at 1–1.5 h after recirculation. (7) Administration of L-NNA (1 mg/kg i.p.) diminished the increments of NO concentration during ischemia and reflow, leading to its neuroprotection. Although it remains to be determined whether NO is neurotoxic or neuroprotective in cerebral ischemia, these results directed us to the conclusion that NOS activation in brain microvessels and cerebral cortex may play a role in the cellular mechanisms underlying ischemic brain injury.
Article
Full-text available
IL-6 has been reported to have neuroprotective effects against cerebral ischemia while IL-8 is a pro inflammatory cytokine structurally related to interleukin-1 family. In the present study, we tried to determine whether 2% Creatine monohydrate supplementation for variable duration influence the IL-6 and 18 concentrations in the serum of male albino mouse following right common carotid artery ligation and hypoxia (8% oxygen) for 25 minutes. Our result revealed that serum concentration of IL6 (P=0.0001) as well as IL-18 (P=0.003) were significantly higher in mice supplemented with creatine monohydrate for 15 weeks than in male albino mice on normal rodent diet following hypoxic ischemic insult indicating that long term creatine monohydrate supplementation up regulates the IL-6 and IL-18 concentrations triggering the neuroinflammatory and neuroprotective responses.
Article
Full-text available
Objective: Compare the prevalence of brain magnetic resonance imaging abnormalities between patients with or without CD; determine if inflammatory biomarkers are increased in CD; and determine the efficacy of aspirin in reducing the rate of microembolization in these patients. Methods: 500 consecutive patients with heart failure will undergo a structured cognitive evaluation, biomarker collection and search for microembolic signals on transcranial Doppler. The first 90 patients are described, evaluated with cognitive tests and brain magnetic resonance imaging to measure N-acetyl aspartate (NAA), choline (Cho), myo-inositol (MI) and creatine (Cr). Results: Mean age was 55±11 years, 51% female, 38 (42%) with CD. Mean NAA/Cr ratio was lower in patients with CD as compared to other cardiomyopathies. Long-term memory and clock-drawing test were also significantly worse in CD patients. In the multivariable analysis correcting for ejection fraction, age, sex and educational level, reduced NAA/Cr (p=0.006) and cognitive dysfunction (long-term memory, p=0.023; clock-drawing test, p=0.015) remained associated with CD. Conclusion: In this preliminary sample, CD was associated with cognitive impairment and decreased NAA/Cr independently of cardiac function or educational level.
Article
Background: The genetic polymorphism that contributes to the development of infantile cerebral palsy (ICP) after a perinatal hypoxia-ischemia episode remains unknown. Because IL-1? plays a crucial role in the pathogenesis of hypoxic-ischemic encephalopathy, we evaluated whether the interleukin 1, beta (IL-1?) promoter single-nucleotide polymorphisms (SNPs) correlate with the increased risk for ICP following a perinatal hypoxia-ischemia. We assessed -511 C>T and -31 T>C IL-1? SNPs known to be involved in IL-1? expression. Methods: Genomic DNAs were purified from peripheral leukocytes of 48 ICP patients and 57 healthy children, amplified by using polymerase chain reaction (PCR), and then analyzed by using the restriction-fragment-length polymorphism technique. The SNP genotypes were established using real time PCR and validated with the restriction enzymes AvaI for -511C-T and AluI for -31T-C in Restriction fragment length polymorphism (RFLP) analysis. Results: Allelic frequencies of the IL-1? -511 T carrier in patients were significantly higher when compared with those determined in healthy controls. The -511 TT genotype frequency showed a significant difference [odds ratio = 2.4 (95% confidence interval 1.7-3.5), P = 0.0001, and relative risk = 1.5 (95% confidence interval 1.3-1.7)] between patients and controls. The SNP frequencies of -31 genotypes in cerebral palsy patients were not statistically different from the healthy controls. Conclusions: Mexican children with the homozygous TT mutation in the -511 SNP of the IL-1? gene promoter are 2.4 times more susceptible to develop ICP after suffering perinatal asphyxia than healthy children and the presence of a single allele C could be considered as a genetic protective factor.
Article
Background and purpose: The significance and physiological implications of the expression of the 72-kd heat-shock protein in ischemic tissue are unknown. To enhance our understanding of the relation between ischemic cell damage and 72-kd heat-shock protein expression, we evaluated the cellular expression and the anatomic distribution of 72-kd heat-shock protein in conjunction with the morphological analysis of rat brain, as a function of the duration of a single arterial occlusion. Methods: Adult Wistar rats were subjected to graded transient middle cerebral artery occlusion (for a duration of 10, 20, 30, 60, 90, and 120 minutes and sham; n = 4 per group). Forty-eight hours after reopening the artery, brain tissue sections were analyzed to determine the extent of neuronal damage (hematoxylin and eosin staining), the extent of astrocytic reactivity (immunohistochemistry, using anti-glial fibrillary acidic protein), and the distribution of 72-kd heat-shock protein (immunohistochemistry, using a monoclonal antibody to 72-kd heat-shock protein). Results: We found that 72-kd heat-shock protein was sequentially expressed in morphologically intact neurons, microglia, and endothelial cells with increasing duration of ischemia; 72-kd heat-shock protein immunoreactivity was not detected in astrocytes. The duration of ischemia required to evoke a 72-kd heat-shock protein response in neurons was dependent on the anatomic site and followed a pattern of increasing neuronal sensitivity to ischemic cell damage with duration of ischemia: 72-kd heat-shock protein and neuronal damage were sequentially detected in the caudate putamen, globus pallidus, cerebral cortex, amygdala, and hippocampus with increasing duration of ischemia. With ischemia of long duration (greater than or equal to 90 minutes), neurons expressing 72-kd heat-shock protein were localized to a zone peripheral to the severely damaged ischemic core. Conclusions: These studies suggest that 1) the expression of 72-kd heat-shock protein in neurons precedes the development of ischemic cellular alterations detectable by conventional hematoxylin and eosin light microscopy methods; 2) there is a hierarchy of cell types and anatomic sites that express 72-kd heat-shock protein, and this hierarchy reflects cellular and anatomic vulnerability to ischemic cell damage; and 3) 72-kd heat-shock protein induction in neurons bordering a necrotic ischemic core may be the morphological equivalent of the ischemic penumbra.
Article
We have examined the incidence and size of infarction after occlusion of different portions of the rat middle cerebral artery (MCA) in order to define the reliability and predictability of this model of brain ischemia. We developed a neurologic examination and have correlated changes in neurologic status with the size and location of areas of infarction. The MCA was surgically occluded at different sites in six groups of normal rats. After 24 hr, rats were evaluated for the extent of neurologic deficits and graded as having severe, moderate, or no deficit using a new examination developed for this model. After rats were sacrificed the incidence of infarction was determined at histologic examination. In a subset of rats, the size of the area of infarction was measured as a percent of the area of a standard coronal section. Focal (1-2 mm) occlusion of the MCA at its origin, at the olfactory tract, or lateral to the inferior cerebral vein produced infarction in 13%, 67%, and 0% of rats, respectively (N = 38) and produced variable neurologic deficits. However, more extensive (3 or 6 mm) occlusion of the MCA beginning proximal to the olfactory tract--thus isolating lenticulostriate end-arteries from the proximal and distal supply--produced infarctions of uniform size, location, and with severe neurologic deficit (Grade 2) in 100% of rats (N = 17). Neurologic deficit correlated significantly with the size of the infarcted area (Grade 2, N = 17, 28 +/- 5% infarction; Grade 1, N = 5, 19 +/- 5%; Grade 0, N = 3, 10 +/- 2%; p less than 0.05). We have characterized precise anatomical sites of the MCA that when surgically occluded reliably produce uniform cerebral infarction in rats, and have developed a neurologic grading system that can be used to evaluate the effects of cerebral ischemia rapidly and accurately. The model will be useful for experimental assessment of new therapies for irreversible cerebral ischemia.
Article
: The expression of interleukin-1β (IL-1β) mRNA in the cerebral cortex, hippocampus, striatum, and thalamus of rats was studied after transient forebrain ischemia. IL-1β mRNA was not detected in all these regions of sham-operated control rats. IL-1β mRNA was induced after transient forebrain ischemia and reached a detectable level in all regions examined 15 min after the start of recirculation. The induction of IL-1β mRNA had a few peaks, that is, peaks were observed at 30 and 240 min in the four regions examined, and another peak was observed at 90 min in the striatum. One day after the start of recirculation, IL-1β mRNA levels were markedly decreased, but even 7 days after that, IL-1β mRNA was found at very low levels in all regions examined. The amounts of c-fos and β-actin mRNAs on the same blots were also examined. The induction of c-fos mRNA was transient and had only one peak in all regions examined, whereas the levels of β-actin mRNA in these regions were fairly constant throughout the recirculation period. Thus, we provide the first evidence for a characteristic expression of IL-1β mRNA in several brain regions after transient forebrain ischemia.
Article
Article
Neutrophils are critically involved with ischemia and reperfusion injury in many tissues but have not been studied under conditions of reperfusion after focal cerebral ischemia. The present studies were conducted to confirm our previous observations quantifying neutrophils in rat permanent focal stroke using a myeloperoxidase activity assay and to extend them to transient ischemia with reperfusion. In addition, leukotriene B4 receptor binding in ischemic tissue was evaluated as a potential marker for inflammatory cell infiltration. Histological, enzymatic, and receptor binding techniques were used to evaluate neutrophil infiltration and receptor binding in infarcted cortical tissue 24 hours after permanent middle cerebral artery occlusion (n = 25) or temporary occlusion for 80 (n = 12) or 160 (n = 22) minutes followed by reperfusion for 24 hours in spontaneously hypertensive rats. Sham surgery (n = 26) produced no changes in any parameter measured. After permanent middle cerebral artery occlusion, neutrophil accumulation was observed histologically, but the infiltration was moderate and typically within and adjacent to blood vessels bordering the infarcted cortex. After temporary middle cerebral artery occlusion with reperfusion, marked neutrophil infiltration was observed throughout the infarcted cortex. Myeloperoxidase activity was increased (p less than 0.05) after permanent occlusion and to a greater extent after temporary occlusion with reperfusion. Myeloperoxidase activity (units per gram wet weight) in ischemic cortex was increased over that in nonischemic (control) cortex 32.2-fold, 54.6-fold, and 92.1-fold for permanent occlusion and 80 and 160 minutes of temporary occlusion with reperfusion, respectively (p less than 0.05). Sham surgery produced no changes in myeloperoxidase activity. Leukotriene B4 receptor binding also was increased (p less than 0.05) after focal ischemia and paralleled the increases in myeloperoxidase activity. Ischemic cortex-specific receptor binding (femtomoles per milligram protein) was 3.87 +/- 0.63 in sham-operated rats and 4.57 +/- 0.98, 8.98 +/- 1.11, and 11.12 +/- 1.63 for rats subjected to permanent occlusion and 80 and 160 minutes of temporary occlusion with reperfusion, respectively (all p less than 0.05 different from sham-operated). Cortical myeloperoxidase activity was significantly correlated with the degree of cortical leukotriene B4 receptor binding (r = 0.66 and r = 0.79 in two different studies, p less than 0.01). These data indicate that neutrophils are involved in focal ischemia and that there is a dramatic accumulation of neutrophils in infarcted tissue during reperfusion that can be quantified using the myeloperoxidase activity assay. Leukotriene B4 receptor binding increases in infarcted tissue in a parallel manner, which suggests that the increased leukotriene B4 binding is to receptors located on the accumulating neutrophils.
Article
The significance and physiological implications of the expression of the 72-kd heat-shock protein in ischemic tissue are unknown. To enhance our understanding of the relation between ischemic cell damage and 72-kd heat-shock protein expression, we evaluated the cellular expression and the anatomic distribution of 72-kd heat-shock protein in conjunction with the morphological analysis of rat brain, as a function of the duration of a single arterial occlusion. Adult Wistar rats were subjected to graded transient middle cerebral artery occlusion (for a duration of 10, 20, 30, 60, 90, and 120 minutes and sham; n = 4 per group). Forty-eight hours after reopening the artery, brain tissue sections were analyzed to determine the extent of neuronal damage (hematoxylin and eosin staining), the extent of astrocytic reactivity (immunohistochemistry, using anti-glial fibrillary acidic protein), and the distribution of 72-kd heat-shock protein (immunohistochemistry, using a monoclonal antibody to 72-kd heat-shock protein). We found that 72-kd heat-shock protein was sequentially expressed in morphologically intact neurons, microglia, and endothelial cells with increasing duration of ischemia; 72-kd heat-shock protein immunoreactivity was not detected in astrocytes. The duration of ischemia required to evoke a 72-kd heat-shock protein response in neurons was dependent on the anatomic site and followed a pattern of increasing neuronal sensitivity to ischemic cell damage with duration of ischemia: 72-kd heat-shock protein and neuronal damage were sequentially detected in the caudate putamen, globus pallidus, cerebral cortex, amygdala, and hippocampus with increasing duration of ischemia. With ischemia of long duration (greater than or equal to 90 minutes), neurons expressing 72-kd heat-shock protein were localized to a zone peripheral to the severely damaged ischemic core. These studies suggest that 1) the expression of 72-kd heat-shock protein in neurons precedes the development of ischemic cellular alterations detectable by conventional hematoxylin and eosin light microscopy methods; 2) there is a hierarchy of cell types and anatomic sites that express 72-kd heat-shock protein, and this hierarchy reflects cellular and anatomic vulnerability to ischemic cell damage; and 3) 72-kd heat-shock protein induction in neurons bordering a necrotic ischemic core may be the morphological equivalent of the ischemic penumbra.