Tải bản đầy đủ (.pdf) (10 trang)

Báo cáo khoa học: Post-ischemic brain damage: targeting PARP-1 within the ischemic neurovascular units as a realistic avenue to stroke treatment pptx

Bạn đang xem bản rút gọn của tài liệu. Xem và tải ngay bản đầy đủ của tài liệu tại đây (207.18 KB, 10 trang )

MINIREVIEW
Post-ischemic brain damage: targeting PARP-1 within the
ischemic neurovascular units as a realistic avenue to
stroke treatment
Flavio Moroni and Alberto Chiarugi
Department of Preclinical and Clinical Pharmacology, University of Florence, Italy
Therapeutic strategies aimed at reducing brain dam-
age after ischemic stroke have been a major focus of
academic and industrial research for the past
30 years. Two primary therapeutic approaches have
been intensively studied: the first can be defined as
the ‘vascular approach’ and its main goal is the rapid
re-opening of occluded blood vessels so that oxygen
and nutrients may return to the ischemic region. The
second may be defined as the ‘cellular approach’ and
is based on the possibility of interfering with the sig-
naling pathways, leading to loss of neurons and dam-
age of other cellular elements present in the affected
brain region [1,2]. Efforts directed at developing effec-
tive vascular therapy led to clinically useful
procedures and have clearly demonstrated that it is
possible to reduce, selectively, brain damage and
neurologic disability by administering recombinant tis-
sue plasminogen activator within 3 h from when the
stroke symptoms first start. Conversely, the cellular
approach has been so far clinically unsuccessful, and
none of the numerous neuroprotective strategies that
have been tested in clinical trials have reached the
clinical arena [3,4].
Exciting, radical, suicidal and
inflamed – the many pathways of


ischemic brain injury
The enormous body of information on ischemic neuro-
degeneration in different experimental stroke models
has shed light on the complex signaling pathways and
molecular events responsible for neuronal damage
Keywords
blood brain barrier; endothelium;
inflammation; ischemia; microglia;
neuroprotection; neurovascular unit;
PARP-1; pericytes; stroke
Correspondence
F. Moroni, Dipartimento di Farmacologia,
Viale Pieraccini 6, 50139 Firenze, Italy
Fax: +39 055 4271226
Tel: +39 055 4271280
E-mail: flavio.moroni@unifi.it
(Received 3 July 2008, revised 11
September 2008, accepted 14 October
2008)
doi:10.1111/j.1742-4658.2008.06768.x
Stroke is the third leading cause of death in industrialized countries but
efficacious stroke treatment is still an unmet need. Preclinical research indi-
cates that different molecules afford protection from ischemic neurodegen-
eration, but all clinical trials conducted so far have inexorably failed.
Critical re-evaluation of experimental data shows that all the components
of the neurovascular unit, such as neurons, glia, endothelia and basal mem-
branes, must be protected during the ischemic insult to obtain substantial
and long-lasting neuroprotection. Here, we propose the nuclear enzyme
poly(ADP-ribose) polymerase (PARP-1) as a key effector of cell death in
the various elements of the neurovascular units, and assert that drugs

inhibiting PARP-1 may therefore represent valuable tools for pharmacolog-
ical treatment of stroke patients.
Abbreviations
AIF, apoptosis-inducing factor; BBB, blood–brain barrier; HMGB1, high-mobility-group protein box 1; IL, interleukin; MMP, matrix
metalloproteinase; NMDA, N-methyl-
D-aspartate; PARG, poly(ADP-ribose) glycohydrolase; PARP, poly(ADP-ribose) polymerase; PARP-1,
poly(ADP-ribose) polymerase 1; ROS, reactive oxygen species; TNF-a, tumor necrosis factor-a.
36 FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS
when blood flow to a brain region drops below a criti-
cal threshold and when it returns because of vessel
re-opening and tissue reperfusion. In the past, particu-
lar attention was directed to derangement of excitatory
amino acid-mediated neurotransmission that became,
for years, the main target for neuroprotection.
Hypoxia ⁄ ischemia increases the concentrations of
extracellular glutamate [5,6] with excessive stimulation
of ionotropic and metabotropic glutamate receptors,
which initiates a chain of events leading to excitotoxic
neuronal death [7,8]. This concept is strongly sup-
ported by the observation that, in a number of in vitro
and in vivo experimental models of ischemia, glutamate
receptor antagonists, acting either on ionotropic
[N-methyl-d-aspartate (NMDA) or Gk alpha-amino-3-
hydroxy-5-methyl-4-isoxazolone propinate] or on group
I metabotropic receptors, are effective neuroprotective
agents [9–13]. Unfortunately, however, none of the
glutamate receptor antagonists tested in clinical trials
showed positive results or had an acceptable benefit ⁄
side effects ratio.
Triggered by the excitotoxic events as well as by

impairment of mitochondrial respiration, a burst of
reactive oxygen species (ROS) and reactive nitrogen
species typically occurs within the ischemic brain tis-
sue. Again, inhibition of radical formation as well as
of radical scavengers provides significant neuroprotec-
tion in animal stroke models. Agents acting as free-
radical scavengers therefore have been repeatedly
proposed as useful drugs for stroke therapy, but most
were rapidly discarded because of cardiovascular toxic-
ity. Recently, however, the spin-trap nitrone NXY-059
from AstraZeneca reached the clinical arena with some
success [14]. The putative neuroprotectant is probably
n-t-butyl hydroxylamine and ⁄ or its parent spin-trap
2-methyl-2-nitrosopropane, produced by hydrolysis of
NXY-059. Unfortunately, the positive outcome of the
first clinical trial was not confirmed in a second clinical
trial, and NXY-059 development was dropped, leaving
widespread scepticism in the field regarding the possi-
bility of obtaining ischemic neuroprotection in humans
[15].
Apoptotic mechanisms also contribute to ischemic
neuronal demise. This suicidal form of neurodegenera-
tion seems to occur mainly in specific types of brain
ischemia, including the global type of brain ischemia.
Also, activation of the apoptotic program typically
occurs in a delayed manner in brain regions present in
the surroundings of the ischemic core (the so-called
‘penumbra’, see below) and is thought to be a key com-
ponent of time-dependent brain infarct evolution [1].
Yet, strategies aimed at inhibiting the several apoptotic

effectors have not been exploited at the clinical level.
Another event widely recognized to be of key patho-
genetic relevance to post-ischemic brain damage is
immune activation of resident glial cells and leukocytes
infiltrating from blood vessels [16,17]. In this regard,
several therapeutic approaches aimed at counteracting
post-ischemic immune activation and infiltration have
been tested in clinical trials. Some, such as the anti-leu-
kocyte adhesion molecules enlimonab and HU23F2G,
proved inefficacious and harmful, respectively. Others,
such as the interleukin (IL)-1 receptor antagonist, pro-
vided inconclusive results. Failure might be a result of
the fact that both protective as well as detrimental
effects of the inflammatory response during ischemic
neurodegeneration have been reported [18].
Critical re-evaluation of drug
development in stroke
Preclinical studies clearly show that it is feasible to
protect the brain from ischemic injury by means of
pharmacological or genetic approaches aimed at tar-
geting the molecular mechanisms involved in ischemic
neurodegeneration. Hence, because there are no appar-
ent reasons why these strategies should not be effective
in humans, it is reasonable to predict that effective
neuroprotective strategies identified at the preclinical
level also reach clinical practice. Then, the question is
why has this not yet happened? An increasing body of
literature is accumulating on this subject, and several
critical points that have been identified are the past
and, unfortunately, present criteria and methodologies

used for drug development in the stroke field [3,4,19].
To summarize, it is now clear that animal models
should closely reproduce the complex cardiovascular
and cerebral pathophysiology of stroke patients, and
neuroprotection should be evaluated on a long-lasting
and functional basis, rather than on an acute and his-
tological basis. Also, careful and rigorous selection of
patients with salvageable tissue [evidenced using mag-
netic resonance imaging as the presence of an area of
hypoperfusion larger than that of altered water diffu-
sion (the latter is an index of necrosis), the so-called
‘Perfusion ⁄ Diffusion (PWI ⁄ DWI) mismatch’] should
be conducted before treating them with an anti-stroke
drug candidate [4]. Finally, the concepts of ‘pleiotypic
drugs’ (i.e. drugs with several mechanisms of action)
and ‘synergistic combinatorial drug therapy’ emerge as
key requisites for efficacious stroke treatment [4].
Indeed, one of the possible reasons for the lack of clin-
ical efficacy of drugs tested in clinical trials for brain
ischemia is their selective mechanism of action. For
instance, glutamate antagonists act exclusively (or pre-
dominantly) on neurons. So, even if neurons are the
F. Moroni and A. Chiarugi PARP-1 and the ischemic neurovascular unit
FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS 37
first cell type to lose their function when blood supply
is insufficient, the other cell types present in the ner-
vous tissue are of the utmost importance to support
neuronal functioning. When capillaries and glia are
damaged, neurons cannot survive in spite of protection
from excitotoxic insults. Similarly, selective blocking of

apoptosis or inflammation within the ischemic tissue
cannot provide protection when the other detrimental
events are unrestricted. As a whole, efficacious stroke
treatment needs concomitant targeting of the various
pathogenetic events actively contributing to neurode-
generation in cells localized within the ischemic
penumbra.
Penumbra and the neurovascular unit
The ischemic brain region may be divided into a zone
in which blood flow is completely absent (‘ischemic
core’) and a peripheral zone in which collateral ves-
sels supply only a fraction of the oxygen and glucose
required for the normal activity of neural cells (‘ische-
mic penumbra’) [2,20]. While all cell types in the core
region undergo typical necrotic features and die form-
ing an infarct zone, the ischemic penumbra may
initially retain its morphological integrity, even if its
functions (i.e. electrical activity, synthetic processes,
bioenergetic functions, etc.) are temporally lost. How-
ever, if sufficient blood flow eventually returns to the
ischemic region within a reasonable time (hours) it is
possible to rescue this area, thus limiting the neuro-
logical damage. It is now clear that in order to
obtain full functional recovery, not only neurons, but
all cell types (i.e. astrocytes, microglia, oligodendro-
cytes, endothelial cells, muscle cells, pericytes) and
structures (mainly basal membranes) present in the
‘penumbra area’ should be rescued [21,22]. Thus,
ischemic neuroprotection can be achieved only if the
classic, oversimplified strategy ‘save the neurons’ is

changed into ‘save neural and stromal cells’. Overall,
neural and stromal cells are grouped into a functional
entity: the so-called ‘neurovascular unit’. Operatively,
the latter is a very complex network of functions
brought about by different cells and aimed at main-
taining the homeostatic milieu necessary for normal
brain activities. Protection of the components of the
neurovascular unit seems therefore essential to reduce
brain damage and neurological deficits after a stroke.
To achieve this, different strategies have been pro-
posed and evaluated in preclinical settings. Yet, con-
comitant targeting of all the components of the
neurovascular units adds substantial complexity to
the feasibility of obtaining ischemic neuroprotection
by pharmacological approaches and, as mentioned
above, general scepticism permeated the field. As
outlined below, we claim that poly(ADP-ribose) poly-
merase 1 (PARP-1) inhibitors are among the most
efficacious protectants of the neurovascular unit
currently available.
PARP-1 activation and cell death in the
neurovascular unit
Poly(ADP-ribose) polymerases (PARPs) are NAD-
dependent enzymes that are able to catalyse the trans-
fer of ADP-ribose units from NAD to substrate
proteins, thereby contributing to the control of geno-
mic integrity, cell cycle and gene expression [23].
Among PARPs, nuclear PARP-1 is a DNA damage-
activated enzyme of 113 kDa molecular mass and is
the most abundant and commonly studied member of

the family. Its enzymatic activity leads to poly(ADP
ribose) formation, and it was first described over
40 years ago in liver cell nuclei incubated with NAD
and ATP in Paul Mandel’s laboratory in Strasburg
[24]. Although this seminal observation was made in a
neuroscience laboratory, for the following 30 years,
research on PARP-1 was exclusively carried out by
researchers mainly involved in genome stability, DNA
repair and cancer. The neuroscience community
ignored PARP-1 until the early 1990s when it was
shown that it mediates glutamate-induced and nitric
oxide-induced neuronal death [25,26]. Excellent work
carried out in the following years uncovered several
molecular events causally linking PARP-1 activation to
ischemic cell death [27]. As for the triggers of PARP-1
hyperactivity during ischemia, ROS-dependent DNA
damage is thought to play a major role. However,
Ca
2+
-dependent and kinase-dependent PARP-1 activa-
tion might also contribute [28–30]. Ambiguity also
exists regarding the molecular mechanisms underlying
the detrimental role of the enzyme in ischemic brain
injury [31,32]. Indeed, although we know in part the
mechanisms activated by PARP-1 and triggering
neurotoxicity, which of these is causally involved in
PARP-1-dependent ischemic neurodegeneration still
needs to be elucidated.
Experimental data demonstrate that, upon different
stresses, activation of PARP-1 can exert detrimental

effects in every cell type of the neurovascular unit
(Fig. 1). Given that the ischemic challenge mimics
these stresses, we reason that during brain ischemia
PARP-1-dependent cytotoxicity occurs in all the com-
ponents of the neurovascular unit. It is obvious that
triggers, time courses and final effects of PARP-1
activation in endothelial, muscle and glial cells, as well
as in infiltrating leukocytes, are different from those
PARP-1 and the ischemic neurovascular unit F. Moroni and A. Chiarugi
38 FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS
occurring in neurons. Regardless, the hyperactivation
of PARP-1 in each single component of the neurovas-
cular unit triggers dysfunction ⁄ cytotoxicity and, indi-
rectly, severely affects the functioning of neighbouring
neurons. As a whole, PARP-1-dependent derangement
of the integrity of the neurovascular unit is caused by
the enzyme’s ability to prompt an increase of blood–
brain barrier (BBB) permeability, the release of pro-
inflammatory mediators, mitochondrial dysfunction
and bioenergetic failure, as well as the activation of
specific apoptotic pathways.
PARP-1, endothelia and post-ischemic
BBB breakdown
Ischemia causes rapid structural changes and break-
down of the BBB, allowing plasma exudation and
immune cell infiltration, which contribute to ischemic
brain damage [22]. Very early after the onset of brain
ischemia, and especially after a reperfusion period,
abundant free radicals are generated in macrophages,
endothelial cells, perycites, astrocytes, microglia and

neurons, causing significant damage to brain capillaries
and disruption of the BBB [33]. Free radicals formed
both inside and outside the vessels prompt genotoxic
stress and activate PARP-1 in endothelial cells. Under
conditions of chronic hypoxia, PARP-1 activation
within endothelia triggers cell proliferation and slowly
developing brain damage. The molecular mechanisms
of cell proliferation include the generation and release
of ROS from NADPH oxidase and mitochondria, sus-
tained increase of the cytosolic Ca
2+
concentration
and finally nuclear translocation of mitogen-activated
protein kinase kinase ⁄ extracellular regulated protein
kinase with cell cycle activation [34]. Conversely,
during ischemia, PARP-1 hyperactivation causes endo-
thelial cell death. The latter occurs because of cellular
accumulation of the PARP-1 product poly(ADP-
ribose), which causes translocation of apoptosis-induc-
ing factor (AIF) from mitochondria to the nucleus and
activation of a caspase-independent programmed cell-
death pathway [35–37]. Accordingly, the potent
PARP-1 inhibitor, PJ34, administered to rats with
transient focal brain ischemia, preserves the integrity
M
M
P
M
M
P

1-
P
RAP
1-PRAP
1
-
PRAP
1-PR
AP
1-
PR
AP
1
-PRAP
FI
A
FI
A
T
R
P
M
2
aC
+
2
1BG
M
H
H

M
G
B
1
y
rot
amm
alfnI
s
r
o
t
a
idem
y
r
ota
mm
alf
n
I
srotaidem
X
X
X
X
M
M
P
yrotammalf

n
I
srotaid
e
m
norueN
M
i
c
r
o
g
l
i
a
etycortsA
et
yc
o
kueL
muilehto
dn
E
nemuL
Basal lamina
Fig. 1. The role of PARP-1 within the ischemic neurovascular units. PARP-1 exerts its detrimental role within the ischemic neurovascular
unit by promoting necrosis and AIF-dependent apoptosis in neurons, astrocytes and endothelial cells. PARP-1 also plays a key role in
immune activation and migration of microglial cells upon different noxious stimuli to the central nervous system. The expression of adhesion
molecules by endothelial cells is also promoted by PARP-1-dependent transcriptional activation, thereby promoting leukocyte recruitment
within the ischemic brain tissue and their detrimental effects on ischemic injury. Hence, the pharmacological inhibition of the enzyme exerts

ischemic neuroprotection by targeting several pleiotypic events of pathogenetic relevance to post-ischemic brain damage. X, adhesion mole-
cules. ADP-ribose monomers are depicted as black circles binding to the transient receptor potential melastatin-2 receptor.
F. Moroni and A. Chiarugi PARP-1 and the ischemic neurovascular unit
FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS 39
of endothelial tight junctions and decreases the expres-
sion of the adhesion molecule intercellular adhesion
molecule-1, thus limiting leukocyte infiltration and the
subsequent inflammatory damage to the ischemic brain
[35,38]. It has also been proposed that post-ischemic
PARP-1 activation contributes to increased expression
of matrix metalloproteinases (MMPs), a group of zinc-
containing proteases with key roles in matrix degrada-
tion and disruption of capillary permeability during
stoke [39]. Indeed, pharmacological PARP-1 inhibition
reduces MMP-9 expression levels in plasma and brain
[40], prevents brain matrix degradation, reduces
delayed increase of BBB permeability and edema for-
mation, preserves endothelial tight junction proteins
and decreases delayed infiltration of leukocytes into
the brain of rats with middle cerebral artery occlusion
[41]. The key role of PARP-1 hyperactivation in endo-
thelial dysfunction in experimental models of diabetes
underscores the pathogenetic relevance of the enzyme
to disorders of this key component of the neurovascu-
lar unit [42]. Accordingly, gene array studies have
demonstrated that upregulation of inflammatory genes
is hampered in PARP-1
) ⁄ )
endothelial cells exposed to
tumor necrosis factor-alfa (TNF-a) [43]. Taken

together, these findings point to basal PARP-1 activity
as central to homeostatic regulation of endothelial
function, whereas its hyperactivation appears causal
to BBB damage and immune cell infiltration during
ischemia.
PARP-1, glia and post-ischemic
inflammatory events
Activation of resident immune cells as well as infiltra-
tion of leukocytes within the ischemic area lead to
excessive release of inflammatory mediators and ensu-
ing worsening of brain damage. In keeping with this,
astrocytes, microglia and blood-derived leukocytes
contribute to ischemic neurodegeneration, whereas
immunosuppressant strategies able to reduce the
inflammatory response decrease infarct volumes in dif-
ferent stroke models [16,17]. Microglial cells are resi-
dent brain macrophages displaying a ‘resting’ highly
ramified phenotype. Upon ischemic challenge, before
neuronal damage can be morphologically detected [44],
microglia assume amoeboid morphology and acquire
phagocytic activity, producing ROS and other inflam-
matory ⁄ cytotoxic factors such as nitric oxide, prosta-
noids, TNF-a, IL-1b and MMPs. Astrocytes and
infiltrating leukocytes within the ischemic brain tissue
also contribute to the synthesis and release of pro-
inflammatory mediators [17]. It is now widely accepted
that the latter are responsible for disruption of the
capillary basal lamina, opening of the BBB and infil-
tration of blood-borne leukocytes. This prompts a
vicious circle comprising waves of release of cytotoxic

inflammatory products, cell death and recruit-
ment ⁄ activation of blood or bystander immune cells.
Eventually, the neuroimmune response causes collapse
of the structures and functions of the neurovascular
unit [16,17,45].
Again, PARP-1 plays a key role in this scenario.
Indeed, numerous reports demonstrate that PARP-1
activity promotes the neuroimmune response thanks
to its ability to assist transcriptional activation and
epigenetic remodeling in immune cells. In this light, it
has been speculated that ischemic neuroprotection
afforded by PARP inhibitors is at least partially med-
iated by their anti-inflammatory properties [46].
Indeed, PARP inhibitors decrease expression of
inflammatory markers ⁄ mediators such as CD11b,
cyclooxygenase-2, inducible nitric oxide synthase,
TNF-a, IL-1b, IL-6, intracellular adhesion molecule-1,
interferon-gamma and E-selectin in different models
of neurodegeneration [40,47–55]. Remarkably, these
molecules actively contribute to ischemic neurodegen-
eration. A key role for PARP-1 in microglia activa-
tion and migration towards injured neurons has also
been reported [56]. Reduced expression of pro-inflam-
matory mediators is probably a result of the fact that
inflammatory transcription factors such as nuclear
factor-kappaB, activator protein-1 and nuclear factor
of activated T-cells are positively regulated by
PARP-1. PARP-1 protein per se , as well as its enzy-
matic activity, promote transcription factor binding
to DNA as well as supramolecular complex formation

containing several transcription-regulating proteins
and RNA polymerase II [23,53,57]. These findings
taken together may explain why post-treatment with
PARP-1 inhibitors reduces the neuroimmune response
in different stroke models [58–60].
Recently, the tetracycline, minocycline, has been
proposed as a clinically relevant tool to limit post-
ischemic brain damage because of its ability to inhibit
microglia activation. Minocycline is indeed able to
reduce brain infarct volumes in preclinical models
[61], as well as neurological impairment in stroke
patients [62]. Interestingly, it has recently been
reported that minocycline is a powerful inhibitor of
PARP-1 [63]. Whether PARP-1 inhibition underpins
the drug’s neuroprotective effects in stroke patients is
currently unknown. Yet, given that minocycline has
been largely used without significant side effects, these
observations indicate that acute inhibition of PARP-1
in vivo might be a rather safe procedure and could be
proposed to preserve the integrity of the ischemic
PARP-1 and the ischemic neurovascular unit F. Moroni and A. Chiarugi
40 FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS
neurovascular unit and limit post-ischemic brain
damage in humans.
PARP-1 and post-ischemic death in
neurons
Excitotoxicity and PARP-1 activation have been caus-
ally linked since 1994 when it was reported that gluta-
mate increases poly(ADP-ribose) synthesis and causes
a type of cell death that is prevented by both NMDA

antagonists and PARP-1 inhibitors [25,26]. The pro-
posed molecular events underlying these observations
include: overactivation of NMDA glutamate receptors
with consequent intracellular Ca
2+
influx; and subse-
quent ROS production mainly caused by neuronal
nitric oxide synthase activity, which, in turn, triggers
DNA damage-dependent hyperactivation of PARP-1,
depletion of intracellular NAD and ATP stores, and
neuronal death [26]. PARP-1 activation may also occur
in neurons without NMDA receptor activation, as
increases of intracellular [Ca
2+
] triggered by K
+
-
induced depolarization or inositol 3-phosphate-recep-
tor activation are sufficient to trigger poly(ADP-ribose)
formation [28,64]. In keeping with this toxic cascade of
events, neurons obtained from PARP-1-deficient mice
are resistant to NMDA toxicity and to oxygen and
glucose deprivation [65]. It was also shown that
NMDA-induced overload of cytosolic Ca
2+
not only
activates neuronal nitric oxide synthase in the cytosol,
but is also responsible for mitochondrial ROS produc-
tion [66], which contributes to DNA damage and fur-
ther activation of PARP-1 [67,68]. Substantial DNA

damage, evaluated by means of the comet assay, is
present in cells isolated from the rat ischemic cortex or
caudate. NMDA receptor antagonists reduce the
extent of the damage and provide ischemic neuropro-
tection, while PARP inhibitors decrease infarct vol-
umes without affecting the severity of DNA damage
[69]. These observations suggest that NMDA receptor
channel openings, ROS formation, DNA damage and
PARP activation are sequential crucial steps in the
process leading to neuronal death. They also indicate
that stroke protection can be achieved without reduc-
ing DNA damage. Energy failure following PARP-1
activation is not only caused by NAD resynthesis but
also by glycolysis block because of NAD depletion,
which results in reduced synthesis of both glycolysis-
derived ATP and mitochondrial energetic substrates
[70]. Accordingly, tricarboxylic acid cycle substrates or
extracellular NAD supplementation protect neurons
from excessive PARP-1 activation [71], whereas
PARP-1 inhibitors prevent ischemia-induced NAD
+
depletion and reduce ischemic brain injury [72]. In
apparent contrast to the hypothesis that PARP-1
worsens ischemic neurodegeneration by reducing ATP
levels within the injured tissue, however, ischemia-
induced energy derangement is similar in the affected
brain areas of PARP-1
+ ⁄ +
and PARP-1
) ⁄ )

mice,
despite the latter showing significant reduction of
ischemic volumes [73].
Controversy still exists on the molecular mecha-
nisms involved in PARP-1-dependent neuronal death
during ischemia. In this regard it has been very
recently reported that exposure of cultured neurons
to poly (ADP-ribose) is sufficient to trigger nuclear
translocation of mitochondrial AIF and cell demise
[74]. The poly(ADP-ribose)-degrading enzyme, poly
(ADP-ribose) glycohydrolase (PARG), should be, in
principle, a neuroprotective agent [75]. Consistently,
PARG-110 kDa
) ⁄ )
or PARG
+ ⁄ )
mice show increased
sensitivity to brain ischemia [36,76]. Also, PARP-1
activity seems to be essential for AIF release within
neurons of the infarct area, and AIF-deficient (Harle-
quin) mice are less sensitive to post-ischemic brain
damage [77]. Data therefore point to PARP-1 activ-
ity-dependent AIF release from mitochondria as a
key molecular event underlying ischemic neuronal
death. Interestingly, the ADP-ribose monomers origi-
nating from the polymer degradation through PARG
might also contribute to neuronal demise by activat-
ing transient receptor potential melastatin-2 receptors
and massive Ca
2+

influx [78,79]. Finally, the finding
that, when released in the extracellular space, high-
mobility-group protein box 1 (HMGB1) promotes the
neuroinflammatory response and worsens brain ische-
mia [80–82], along with evidence that PARP-1 pro-
motes HMGB1 release [83] (but also see [82]),
indicate that HMGB1 may mediate, in part, the toxic
effect of PARP-1 hyperactivation within the ischemic
brain tissue. Overall, a wealth of evidence points to
the synthesis of poly (ADP-ribose) within ischemic
neurons as a crucial event contributing to derange-
ment of the neurovascular unit.
Conclusion
To reduce brain damage after stroke it is not sufficient
to protect neurons from excitotoxic insults, but it is
mandatory to rescue all cellular and structural compo-
nents of the neurovascular unit. As outlined above,
PARP-1 activation during brain ischemia plays a detri-
mental role in all cell types of the neurovascular unit.
Inhibitors of PARP-1 might therefore represent a class
of ‘pleiotypic drugs’, which are considered the most
promising tools for pharmacological treatment of
stroke. Also, the different temporal kinetics of PARP-1
F. Moroni and A. Chiarugi PARP-1 and the ischemic neurovascular unit
FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS 41
activation within the components of the neurovascular
unit would warrant a significant ‘window of opportu-
nity’ to be harnessed for the treatment of stroke
patients. Remarkably, the clinical relevance of PARP-1
inhibitors in stroke treatment is emphasized by the fact

that these drugs are well tolerated by patients enrolled
in clinical trials for treatment of tumor malignancies
or coronary bypass, and that, theoretically, anti-stroke
treatment with PARP-1 inhibitors would require an
acute, 4–6-day treatment. This, of course, would
reduce the risk of side effects. The latter might be fur-
ther reduced by the forthcoming development of
PARP isoform-specific inhibitors [84]. In conclusion,
preclinical and clinical data indicate that PARP-1 is a
very promising target for ischemic neuroprotection,
and PARP-1 inhibitors represent a realistic new avenue
to stroke treatment.
References
1 Lo EH, Dalkara T & Moskowitz MA (2003) Mecha-
nisms, challenges and opportunities in stroke. Nat Rev
Neurosci 4, 399–415.
2 Dirnagl U, Iadecola C & Moskowitz MA (1999) Patho-
biology of ischaemic stroke: an integrated view. Trends
Neurosci 22, 391–397.
3 O’Collins VE, Macleod MR, Donnan GA, Horky LL,
van der Worp BH & Howells DW (2006) 1,026 experi-
mental treatments in acute stroke. Ann Neurol 59,
467–477.
4 Savitz SI & Fisher M (2007) Future of neuroprotection
for acute stroke: in the aftermath of the SAINT trials.
Ann Neurol 61, 396–402.
5 Benveniste H, Drejer J, Schousboe A & Diemer N
(1984) Elevation of extracellular concentrations of glu-
tamate and aspartate in rat hippocampus during tran-
sient cerebral ischemia monitored by intracerebral

microdialysis. J Neurochem 43, 1369–1374.
6 Lombardi G & Moroni F (1992) GM1 ganglioside
reduces ischemia-induced excitatory amino acid output:
a microdialysis study in the gerbil hippocampus. Neuro-
sci Lett 134, 171–174.
7 Choi DW (1992) Excitotoxic cell death. J Neurobiol 23,
1261–1276.
8 Lipton SA & Rosenberg PA (1994) Excitatory amino
acids as a final common pathway for neurological disor-
ders. N Engl J Med 330, 613–622.
9 Meldrum B & Garthwaite J (1990) Excitatory amino
acid neurotoxicity and neurodegenerative disease.
Trends Pharmacol Sci 11, 379–387.
10 Nicoletti F, Bruno VM, Copani A, Casabona G &
Knopfel T (1996) Metabotropic glutamate receptors: a
new target for the therapy of neurodegenerative disor-
ders? Trends Neurosci 19, 267–271.
11 Moroni F, Lombardi G, Thomsen C, Leonardi P,
Attucci S, Peruginelli F, Albani Torregrossa S, Pelleg-
rini-Giampietro DE, Luneia R & Pellicciari R (1997)
Pharmacological characterization of 1-aminoindan-1,5-
dicarboxylic acid (AIDA), a potent mGluR1 antagonist.
J Pharmacol Exp Ther 281, 721–729.
12 Pellegrini-Giampietro DE, Peruginelli F, Meli E, Cozzi
A, Albani Torregrossa S, Pellicciari R & Moroni F
(1999) Protection with metabotropic glutamate 1
receptor antagonists in models of ischemic neuronal
death: time course and mechanisms. Neuropharmacology
38, 1607–1621.
13 Moroni F, Attucci S, Cozzi A, Meli E, Picca R, Schei-

deler MA, Pellicciari R, Noe C, Sarichelou I & Pelleg-
rini-Giampietro DE (2002) The novel and systemically
active metabotropic glutamate 1 (mGlu1) receptor
antagonist 3-MATIDA reduces post-ischemic neuronal
death. Neuropharmacology 42, 741–751.
14 Lees KR, Zivin JA, Ashwood T, Davalos A, Davis SM,
Diener HC, Grotta J, Lyden P, Shuaib A, Hardemark
HG et al. (2006) NXY-059 for acute ischemic stroke.
N Engl J Med 354, 588–600.
15 Garber K (2007) Stroke treatment–light at the end of
the tunnel? Nat Biotechnol 25, 838–840.
16 Barone FC & Feuerstain GZ (1999) Inflammatory
mediators and stroke: novel opportunities for novel
therapeutics. J Cereb Blood Flow Metab 19, 819–834.
17 Iadecola C & Alexander M (2001) Cerebral ischemia
and inflammation. Curr Opin Neurol 14 , 89–94.
18 Wang X & Feuerstein GZ (2004) The Janus face of
inflammation in ischemic brain injury. Acta Neurochir
Suppl 89, 49–54.
19 Feuerstein GZ, Zaleska MM, Krams M, Wang X,
Day M, Rutkowski JL, Finklestein SP, Pangalos MN,
Poole M, Stiles GL et al. (2008) Missing steps in the
STAIR case: a Translational Medicine perspective on
the development of NXY-059 for treatment of acute
ischemic stroke. J Cereb Blood Flow Metab 28, 217–
219.
20 Astrup J, Siesjo
¨
BK & Symon L (1981) Thresholds in
cerebral ischemia: the ischemic penumbra. Stroke 12,

723–725.
21 Iadecola C, Goldman SS, Harder DR, Heistad DD,
Katusic ZS, Moskowitz MA, Simard JM, Sloan MA,
Traystman RJ & Velletri PA (2006) Recommendations
of the National Heart, Lung, and Blood Institute work-
ing group on cerebrovascular biology and disease.
Stroke 37, 1578–1581.
22 del Zoppo GJ & Mabuchi T (2003) Cerebral microves-
sel responses to focal ischemia. J Cereb Blood Flow
Metab 23, 879–894.
23 Hassa PO, Haenni SS, Elser M & Hottiger MO (2006)
Nuclear ADP-ribosylation reactions in mammalian
cells: where are we today and where are we going?
Microbiol Mol Biol Rev 70, 789–829.
PARP-1 and the ischemic neurovascular unit F. Moroni and A. Chiarugi
42 FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS
24 Chambon P, Weill JD & Mandel P (1963) Nicotinamide
mononucleotide activation of new DNA-dependent
polyadenylic acid synthesizing nuclear enzyme. Biochem
Biophys Res Commun 11, 39–43.
25 Cosi C, Suzuki H, Milani D, Facci L, Menegazzi M,
Vantini G, Kanai Y & Skaper SD (1994) Poly(ADP-
ribose)polymerase: early involvement in glutamate-
induced neurotoxicity in cultured cerebellar granule
cells. J Neurosci Res 39, 38–46.
26 Zhang J, Dawson VL, Dawson T & Snyder SH (1994)
Nitric oxide activation of poly(ADP-ribose) synthetase
in neurotoxicity. Science 263, 687–689.
27 Szabo C & Dawson VL (1998) Role of poly(ADP-
ribose) synthetase in inflammation and ischaemia-

reperfusion. Trends Pharmacol Sci 19, 287–298.
28 Homburg S, Visochek L, Moran N, Dantzer F, Priel E,
Asculai E, Schwartz D, Rotter V, dekel N & Cohen-
Armon M (2000) A fast signal induced activation of
poly(ADP-ribose)polymerase: a downstream target of
phospholipase C. J Cell Biol 150, 293–307.
29 Szabo C, Pacher P & Swanson RA (2006) Novel modu-
lators of poly(ADP-ribose) polymerase. Trends Pharma-
col Sci 27, 626–630.
30 Alano CC & Swanson RA (2006) Players in the PARP-1
cell-death pathway: JNK1 joins the cast. Trends
Biochem Sci 31, 309–311.
31 Chiarugi A (2005) Poly(ADP-ribosyl)ation and stroke.
Pharmacol Res 52, 15–24.
32 Chiarugi A (2005) Intrinsic mechanisms of poly(ADP-
ribose) neurotoxicity: three hypotheses. Neurotoxicology
26, 847–855.
33 Gursoy-Ozdemir Y, Can A & Dalkara T (2004) Reperfu-
sion-induced oxidative ⁄ nitrative injury to neurovascular
unit after focal cerebral ischemia. Stroke 35, 1449–1453.
34 Abdallah Y, Gligorievski D, Kasseckert SA, Dieterich
L, Schafer M, Kuhlmann CR, Noll T, Sauer H, Piper
HM & Schafer C (2007) The role of poly(ADP-ribose)
polymerase (PARP) in the autonomous proliferative
response of endothelial cells to hypoxia. Cardiovasc Res
73, 568–574.
35 Zhang Y, Zhang X, Park TS & Gidday JM (2005)
Cerebral endothelial cell apoptosis after ischemia-reper-
fusion: role of PARP activation and AIF translocation.
J Cereb Blood Flow Metab 25, 868–877.

36 Yu SW, Andrabi SA, Wang H, Kim NS, Poirier GG,
Dawson TM & Dawson VL (2006) Apoptosis-inducing
factor mediates poly(ADP-ribose) (PAR) polymer-
induced cell death. Proc Natl Acad Sci U S A 103,
18314–18319.
37 Moubarak RS, Yuste VJ, Artus C, Bouharrour A,
Greer PA, Menissier-de Murcia J & Susin SA (2007)
Sequential activation of poly(ADP-ribose) polymerase
1, calpains, and Bax is essential in apoptosis-inducing
factor-mediated programmed necrosis. Mol Cell Biol 27,
4844–4862.
38 Zhang Y, Park TS & Gidday JM (2007) Hypoxic pre-
conditioning protects human brain endothelium from
ischemic apoptosis by Akt-dependent survivin activa-
tion. Am J Physiol Heart Circ Physiol 292, H2573–
H2581.
39 Rosell A & Lo EH (2008) Multiphasic roles for matrix
metalloproteinases after stroke. Curr Opin Pharmacol 8,
82–89.
40 Koh SH, Chang DI, Kim HT, Kim J, Kim MH, Kim
KS, Bae I, Kim H, Kim DW & Kim SH (2005) Effect
of 3-aminobenzamide, PARP inhibitor, on matrix me-
talloproteinase-9 level in plasma and brain of ischemic
stroke model. Toxicology 214, 131–139.
41 Lenzser G, Kis B, Snipes JA, Gaspar T, Sandor P,
Komjati K, Szabo C & Busija DW (2007) Contribution
of poly(ADP-ribose) polymerase to postischemic
blood-brain barrier damage in rats. J Cereb Blood Flow
Metab
27, 1318–1326.

42 Pacher P, Liaudet L, Oriano FG, Abley JG, Czabo E &
Czabo C (2002) The role of poly(ADP-ribose) polymer-
ase activation in the development of myocardial and
endothelial dysfunction in diabetes. Diabetes 51, 514–
521.
43 Carrillo A, Monreal Y, Ramirez P, Marin L, Parrilla P,
Oliver FJ & Yelamos J (2004) Transcription regulation
of TNF-alpha-early response genes by poly(ADP-
ribose) polymerase-1 in murine heart endothelial cells.
Nucleic Acids Res 32, 757–766.
44 Jorgensen MB, Finsen BR, Jensen MB, Castellano B,
Diemer NH & Zimmer J (1993) Microglial and astrogli-
al reactions to ischemic and kainic acid-induced lesions
of the adult rat hippocampus. Exp Neurol 120, 70–88.
45 del Zoppo GJ (2006) Stroke and neurovascular protec-
tion. N Engl J Med 354, 553–555.
46 Chiarugi A (2002) PARP-1: killer or conspirator? The
suicide hypothesis revisited. Trends Pharmacol Sci 23,
122–129.
47 Scott GS, Hake P, Kean RB, Virag L, Szabo C & Hoo-
per DC (2001) Role of poly(ADP-ribose) synthetase
activation in the development of experimental allergic
encephalomyelitis. J Neuroimm 117, 78–86.
48 Chiarugi A (2002) Inhibitors of poly(ADP-ribose) poly-
merase-1 suppress transcriptional activation in lympho-
cytes and ameliorate autoimmune encephalomyelitis in
rats. Br J Pharmacol 137, 761–770.
49 Ha HC, Hester LD & Snyder SH (2002) Poly(ADP-
ribose) polymerase-1 dependence of stress-induced tran-
scription factors and associated gene expression in glia.

Proc Natl Acad Sci U S A 99, 3270–3275.
50 Ha HC, Hester LD & Snyder SH (2002) Poly(ADP-
ribose) polymerase-1 dependence of stress-induced
transcription factors and associated gene expression in
glia. Proc Natl Acad Sci U S A 99, 3270–3275.
51 Chiarugi A & Moskowitz MA (2003) Poly(ADP-ribose)
polymerase-1 activity promotes NF-kappaB-driven tran-
F. Moroni and A. Chiarugi PARP-1 and the ischemic neurovascular unit
FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS 43
scription and microglial activation: implication for neu-
rodegenerative disorders. J Neurochem 85, 306–317.
52 Koh SH, Park Y, Song CW, Kim JG, Kim K, Kim J,
Kim MH, Lee SR, Kim DW, Yu HJ et al. (2004) The
effect of PARP inhibitor on ischaemic cell death, its
related inflammation and survival signals. Eur J Neuro-
sci 20, 1461–1472.
53 Nakajima H, Nagaso H, Kakui N, Ishikawa M, Hira-
numa T & Hoshiko S (2004) Critical role of the auto-
modification of poly(ADP-ribose) polymerase-1 in
nuclear factor-kappaB-dependent gene expression in pri-
mary cultured mouse glial cells. J Biol Chem 279,
42774–42786.
54 Haddad M, Rhinn H, Bloquel C, Coqueran B, Szabo
C, Plotkine M, Scherman D & Margaill I (2006) Anti-
inflammatory effects of PJ34, a poly(ADP-ribose) poly-
merase inhibitor, in transient focal cerebral ischemia in
mice. Br J Pharmacol 149, 23–30.
55 Lee JH, Park SY, Shin HK, Kim CD, Lee WS & Hong
KW (2007) Poly(ADP-ribose) polymerase inhibition by
cilostazol is implicated in the neuroprotective effect

against focal cerebral ischemic infarct in rat. Brain Res
1152, 182–190.
56 Ullrich O, Diestel A, Eyupoglu IY & Nitsch R (2001)
Regulation of microglial expression of integrins by poly
(ADP-ribose) polymerase-1. Nat Cell Biol 3, 1035–1042.
57 Kraus WL (2008) Transcriptional control by PARP-1:
chromatin modulation, enhancer-binding, coregulation,
and insulation. Curr Opin Cell Biol 20, 294–302.
58 Ducrocq S, Benjelloun N, Plotkine M, Ben-Ari Y &
Charriaut-Marlangue C (2000) Poly(ADP-ribose) Syn-
thethase inhibition reduces ischemic injury and inflamma-
tion in neonatal rat brain. J Neurochem 74, 2504–2511.
59 Strosznajder RP, Jesko H & Zambrzycka A (2005)
Poly(ADP-ribose) polymerase: the nuclear target in sig-
nal transduction and its role in brain ischemia-reperfu-
sion injury. Mol Neurobiol 31, 149–167.
60 Hamby AM, Suh SW, Kauppinen TM & Swanson RA
(2007) Use of a poly(ADP-ribose) polymerase inhibitor
to suppress inflammation and neuronal death after cere-
bral ischemia-reperfusion. Stroke 38, 632–636.
61 Yrjanheikki J, Keinanen R, Pellikka M, Hokfelt T &
Koistinaho J (1998) Tetracyclines inhibit microglial acti-
vation and are neuroprotective in global brain ischemia.
Proc Natl Acad Sci U S A 95, 15769–15774.
62 Lampl Y, Boaz M, Gilad R, Lorberboym M, Dabby R,
Rapoport A, Anca-Hershkowitz M & Sadeh M (2007)
Minocycline treatment in acute stroke: an open-label,
evaluator-blinded study. Neurology 69, 1404–1410.
63 Alano CC, Kauppinen TM, Valls AV & Swanson RA
(2006) Minocycline inhibits poly(ADP-ribose) polymer-

ase-1 at nanomolar concentrations. Proc Natl Acad Sci
USA103, 9685–9690.
64 Meli E, Baronti R, Pangallo M, Picca R, Moroni F &
Pellegrini-Giampietro DE (2005) Group I metabotropic
glutamate receptors stimulate the activity of poly(ADP-
ribose) polymerase in mammalian mGlu1-transfected
cells and in cortical cell cultures. Neuropharmacology
49(Suppl 1), 80–88.
65 Pieper A, Verma A, Zhang J & Snyder SH (1999) Poly
(ADP-ribose)polymerase, nitric oxide and cell death.
Trends Pharmacol Sci 20, 171–181.
66 Reynolds IJ & Hastings TG (1995) Glutamate induces
the production of reactive oxygen species in cultured
forebrain neurons following NMDA receptor activa-
tion. J Neurosci 15, 3318–3327.
67 Mandir AS, Poitras MF, Berliner AR, Herring WJ,
Guastella DB, Feldman A, Poirier GG, Wang ZQ,
Dawson TM & Dawson VL (2000) NMDA but not
non-NMDA excitotoxicity is mediated by Poly(ADP-
ribose) polymerase. J Neurosci 20, 8005–8011.
68 Duan Y, Gross RA & Sheu SS (2007) Ca2 + -depen-
dent generation of mitochondrial reactive oxygen spe-
cies serves as a signal for poly(ADP-ribose) polymerase-
1 activation during glutamate excitotoxicity. J Physiol
585, 741–758.
69 Giovannelli L, Cozzi A, Guarnieri I, Dolara P & Mor-
oni F (2002) Comet assay as a novel approach for
studying DNA damage in focal cerebral ischemia: dif-
ferential effects of NMDA receptor antagonists and
poly(ADP-ribose) polymerase inhibitors. J Cereb Blood

Flow Metab 22, 697–704.
70 Ying W, Alano CC, Garnier P & Swanson RA (2005)
NAD+ as a metabolic link between DNA damage and
cell death. J Neurosci Res 79, 216–223.
71 Ying W, Chen Y, Alano CC & Swanson RA (2002) Tri-
carboxylic acid cycle substrates prevent PARP-mediated
death of neurons and astrocytes. J Cereb Blood Flow
Metab 22, 774–779.
72 Endres M, Wang ZQ, Namura S, Waeber C & Mosko-
witz MA (1997) Ischemic brain injury is mediated by
the activation of poly(ADP-ribose)polymerase. J Cereb
Blood Flow Metab 17, 1143–1151.
73 Goto S, Xue R, Sugo N, Sawada M, Blizzard KK, Poi-
tras MF, Johns DC, Dawson TM, Dawson VL, Crain
BJ et al. (2002) Poly(ADP-ribose) polymerase impairs
early and long-term experimental stroke recovery.
Stroke 33, 1101–1106.
74 Andrabi SA, Kim NS, Yu SW, Wang H, Koh DW,
Sasaki M, Klaus JA, Otsuka T, Zhang Z, Koehler RC
et al. (2006) Poly(ADP-ribose) (PAR) polymer is a
death signal. Proc Natl Acad Sci U S A 103, 18308–
18313.
75 Koh DW, Dawson VL & Dawson TM (2005) The
road to survival goes through PARG. Cell Cycle 4,
397–399.
76 Cozzi A, Cipriani G, Fossati S, Faraco G, Formentini
L, Min W, Cortes U, Wang ZQ, Moroni F & Chiarugi
A (2006) Poly(ADP-ribose) accumulation and enhance-
ment of postischemic brain damage in 110-kDa poly
PARP-1 and the ischemic neurovascular unit F. Moroni and A. Chiarugi

44 FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS
(ADP-ribose) glycohydrolase null mice. J Cereb Blood
Flow Metab 26, 684–695.
77 Culmsee C, Zhu C, Landshamer S, Becattini B, Wagner
E, Pellechia M, Blomgren K & Plesnila N (2005) Apop-
tosis-inducing factor triggered by poly(ADP-Ribose)
polymerase and bid mediates neuronal cell death after
oxygen-glucose deprivation and focal cerebral ischemia.
J Neurosci 25, 10262–10272.
78 Fonfria E, Marshall IC, Benham CD, Boyfield I, Brown
JD, Hill K, Hughes JP, Skaper SD & McNulty S (2004)
TRPM2 channel opening in response to oxidative stress
is dependent on activation of poly(ADP-ribose) poly-
merase. Br J Pharmacol 143, 186–192.
79 McNulty S & Fonfria E (2005) The role of TRPM
channels in cell death. Pflugers Arch 451, 235–242.
80 Kim JB, Sig CJ, Yu YM, Nam K, Piao CS, Kim SW,
Lee MH, Han PL, Park JS & Lee JK (2006) HMGB1,
a novel cytokine-like mediator linking acute neuronal
death and delayed neuroinflammation in the post-
ischemic brain. J Neurosci 26, 6413–6421.
81 Faraco G, Fossati S, Bianchi ME, Patrone M, Pedrazzi
M, Sparatore B, Moroni F & Chiarugi A (2007) High
mobility group box 1 protein is released by neural cells
upon different stresses and worsens ischemic neuro-
degeneration in vitro and in vivo. J Neurochem 103,
590–603.
82 Qiu J, Nishimura M, Wang Y, Sims JR, Qiu S, Savitz
SI, Salomone S & Moskowitz MA (2008) Early release
of HMGB-1 from neurons after the onset of brain

ischemia. J Cereb Blood Flow Metab 28, 927–938.
83 Ditsworth D, Zong WX & Thompson CB (2007) Activa-
tion of poly(ADP)-ribose polymerase (PARP-1) induces
release of the pro-inflammatory mediator HMGB1 from
the nucleus. J Biol Chem 282, 17845–17854.
84 Pellicciari R, Camaioni E, Costantino G, Formentini L,
Sabbatini P, Venturoni F, Eren G, Bellocchi D, Chia-
rugi A & Moroni F (2008) On the way to selective
PARP-2 inhibitors. Design, synthesis, and preliminary
evaluation of a series of isoquinolinone derivatives.
ChemMedChem 3, 914–923.
F. Moroni and A. Chiarugi PARP-1 and the ischemic neurovascular unit
FEBS Journal 276 (2009) 36–45 ª 2008 The Authors Journal compilation ª 2008 FEBS 45

×