ArticlePDF AvailableLiterature Review

PPAR: A new pharmacological target for neuroprotection in stroke and neurodegenerative diseases

Authors:

Abstract and Figures

PPARs (peroxisome-proliferator-activated receptors) are ligand-activated transcriptional factor receptors belonging to the so-called nuclear receptor family. The three isoforms of PPAR (alpha, beta/delta and gamma) are involved in regulation of lipid or glucose metabolism. Beyond metabolic effects, PPARalpha and PPARgamma activation also induces anti-inflammatory and antioxidant effects in different organs. These pleiotropic effects explain why PPARalpha or PPARgamma activation has been tested as a neuroprotective agent in cerebral ischaemia. Fibrates and other non-fibrate PPARalpha activators as well as thiazolidinediones and other non-thiazolidinedione PPARgamma agonists have been demonstrated to induce both preventive and acute neuroprotection. This neuroprotective effect involves both cerebral and vascular mechanisms. PPAR activation induces a decrease in neuronal death by prevention of oxidative or inflammatory mechanisms implicated in cerebral injury. PPARalpha activation induces also a vascular protection as demonstrated by prevention of post-ischaemic endothelial dysfunction. These vascular effects result from a decrease in oxidative stress and prevention of adhesion proteins, such as vascular cell adhesion molecule 1 or intercellular cell-adhesion molecule 1. Moreover, PPAR activation might be able to induce neurorepair and endothelium regeneration. Beyond neuroprotection in cerebral ischaemia, PPARs are also pertinent pharmacological targets to induce neuroprotection in chronic neurodegenerative diseases.
Content may be subject to copyright.
International Symposium on Neurodegeneration and Neuroprotection 1341
PPAR: a new pharmacological target for
neuroprotection in stroke and
neurodegenerative diseases
R. Bordet*
1
, T. Ouk*, O. Petrault*, P. Gel
´
e*, S. Gautier*, M. Laprais*, D. Deplanque*, P. Duriez†, B. Staels†,
J.C. Fruchart† and M. Bastide*
*EA1046 Department of Medical Pharmacology, Faculty of Medicine, Institute of Predictive Medicine and Therapeutic Research, University Lille 2 and Lille
University Hospital, 1 place de Verdun, 59045 Lille Cedex, France, and Department of Atherosclerosis, INSERM U545, Institut Pasteur de Lille, 1 rue du Pr.
Calmette B.P. 245, 59019 Lille Cedex, France
Abstract
PPARs (peroxisome-proliferator-activated receptors) are ligand-activated transcriptional factor receptors
belonging to the so-called nuclear receptor family. The three isoforms of PPAR (α, β/δ and γ )areinvolved
in regulation of lipid or glucose metabolism. Beyond metabolic effects, PPARα and PPARγ activation also
induces anti-inflammatory and antioxidant effects in different organs. These pleiotropic effects explain
why PPARα or PPARγ activation has been tested as a neuroprotective agent in cerebral ischaemia. Fibrates
and other non-fibrate PPARα activators as well as thiazolidinediones and other non-thiazolidinedione PPARγ
agonists have been demonstrated to induce both preventive and acute neuroprotection. This neuroprotective
effect involves both cerebral and vascular mechanisms. PPAR activation induces a decrease in neuronal death
by prevention of oxidative or inflammatory mechanisms implicated in cerebral injury. PPARα activation
induces also a vascular protection as demonstrated by prevention of post-ischaemic endothelial dysfunction.
These vascular effects result from a decrease in oxidative stress and prevention of adhesion proteins, such
as vascular cell adhesion molecule 1 or intercellular cell-adhesion molecule 1. Moreover, PPAR activation
might be able to induce neurorepair and endothelium regeneration. Beyond neuroprotection in cerebral
ischaemia, PPARs are also pertinent pharmacological targets to induce neuroprotection in chronic
neurodegenerative diseases.
The treatment of ischaemic stroke is limited to the prevention
of cerebrovascular risk factors and to the modulation of the
coagulation cascade during the acute phase. During the last
two decades, many drugs have been developed to induce
neuroprotection during stroke [1,2]. Nevertheless, none of
them has been successful at the clinical step of their devel-
opment, while recent results give hope of a new antioxidant
drug [3]. One of the explanations for this failure is that the
developed drugs are able to modulate only one molecular
pathway, while several pathways are involved spatially and
temporally in the pathophysiology of stroke. One of the
keys to success in inducing neuroprotection in stroke could
be to modulate simultaneously many pathophysiological
pathways with a combination of several drugs or, better,
with only one pharmacological agent with pleiotropic effect.
Such a pleiotropic effect can be induced by drugs acting on
transcription factor receptors (so-called nuclear receptors),
because this subtype of receptor is able to regulate sev-
eral genes simultaneously. Among nuclear receptors, PPAR
(peroxisome-proliferator-activated receptors) have been
Key words: cerebral ischaemia, neurodegenerative disease, neuroprotection, nuclear receptor,
peroxisome-proliferator-activated receptor (PPAR), thiazolidinedione.
Abbreviations used: Aβ,amyloidβ-peptide; CNS, central nervous system; COX, cyclo-
oxygenase; EAE, encephalomyelitis; MPTP, 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine; NF-
κB, nuclear factor κB; PPAR, peroxisome-proliferator-activated receptor.
1
To whom correspondence should be addressed (email bordet@univ-lille2.fr).
demonstrated to induce pleiotropic effects in different organs
(vessels, heart and kidney) when activated by agonists. Some
results support the idea that these pleiotropic effects could be
useful to induce a neuroprotective effect in stroke [4,5].
PPARs: function and pharmacology
PPARs are ligand-activated transcription factors belonging
to the nuclear receptor superfamily [6] (Figure 1). Three
isoforms of PPARs (α, β/δ and γ ) have been identified, dis-
playing distinct physiological and pharmacological functions
depending on their target genes and their tissue distribution
[7,8]. Indeed, the activation of PPARα, by both natural
ligands such as fatty acids and eicosanoid derivates or
synthetic ligands (lipid-lowering fibrates), regulates lipid
and lipoprotein metabolism [4] (Figure 2). Activation of
PPARγ by prostaglandins or by synthetic ligands such as
antidiabetic thiazolidinediones regulates glucose m etabolism
by modulation of insulin-sensitivity [4]. Non-steroidal anti-
inflammatory drugs are also weak agonists of PPARγ and
PPARα.PPARβ/ δ is one of the most widely expressed
members of the PPAR family. Until recently, the function
of PPARβ/δ remained elusive, but recent results have shown
that PPARβ/δ plays also a key role in lipid metabolism, as
it regulates serum lipid profiles and fatty acid β-oxidation in
muscle and adipose tissue. Synthetic ligands of PPAR β/δ are
at the moment in preclinical phases of development [9].
C
2006 Biochemical Society
1342 Biochemical Society Transactions (2006) Volume 34, part 6
Figure 1 Classical structure of PPAR with the zinc fingers which interact with specific response elements located in DNA
Figure 2 Main physiological effects and pharmacological
modulation of PPAR
PPARs as regulators of inflammation
and oxidative stress
Beyond effects on metabolic pathways, PPAR are also able to
regulate inflammatory pathway by transrepression of trans-
cription factors [NF-κB (nuclear factor κB)] or to regulate the
oxidative pathway [5,6] (Figure 3). PPARα activation induces
expression and activation of antioxidant enzymes such as
superoxide dismutase and glutathione peroxidase. On the
inflammatory pathway, PPARα activation prevents synthesis
and release of cytokines (interleukin-6 and tumour necrosis
factor α) or induction of some inflammatory mediators
such as COX-2 (cyclo-oxygenase 2) and adhesion proteins.
PPARγ activation also reduces the expression of inducible
nitric oxide synthase or COX-2 as well as the production of
pro-inflammatory cytokines. These effects on inflammation
explain why activation of PPAR by synthetic ligands reduces
inflammation in different tissues and in different animal
models of inflammatory diseases (vascular inflammation of
atherosclerosis, inflammatory bowel disease, arthritis etc.)
[4,5].
PPAR in the brain: a potential target
against neuronal death
Previously, it has been supposed that PPAR activation
could also be effective in the regulation of neuronal death
in ischaemic, neurodegenerative or inflammatory cerebral
diseases. Firstly, PPARs have been described in brain and
in spinal cord [10,11]. Beyond expression in cerebral or
spinal blood vessels, PPARs are also expressed in neurons
and in astrocytes, whereas oligodendrocytes exclusively
show PPARβ/δ expression (Figure 4). The extent of this
expression depends on the isoform of PPAR involved.
PPARβ/δ has been found in numerous brain areas, while
PPARα and PPARγ have been localized to more restricted
brains areas [11]. Secondly, whatever the aetiology, neuronal
death is induced by inflammatory and oxidative processes
with a link between the two phenomena [12]. Inflammation
and oxidative stress induce both necrotic and apoptotic
neuronal death. The transcription factor NF-κB plays a
key role in regulation of inflammation and oxidative stress
leading to neuronal death, explaining why PPARs have
been considered as possible targets for neuroprotection [13].
In vitro studies have demonstrated that PPARγ agonists
modulate inflammatory responses to bacterial endotoxin in
brain and also prevent endotoxin-induced neuronal death
[14]. PPARγ agonists are able to prevent neuronal death res-
ulting from NMDA (N-methyl-
D-aspartate) excitotoxicity
induced in brain in vitro or in vivo [15]. PPARα and PPARγ
are able to inhibit macrophage and microglial activation that
C
2006 Biochemical Society
International Symposium on Neurodegeneration and Neuroprotection 1343
Figure 3 Inflammation regulation is a common effect of PPAR and results from modulation of NF-κB
PPAR-responsive element; RXR, retinoid X receptor.
Figure 4 Expression of PPAR in the three cellular t ypes of the neuro-glio-vascular unit
contribute to many degenerative, ischaemic or inflammatory
processes leading to neuronal death [16]. Troglitazone and
ciglitazone inhibit both post-glutamate- and low-potassium-
induced neurotoxicity in cerebellar granule neurons [17].
PPARs are also able to inhibit t he entry of inflammatory
cells into the CNS (central nervous system) from the
periphery by inhibition of chemokines, adhesion molecules
and metalloproteinases [16].
PPAR and cerebral ischaemia
PPAR-induced neuroprotection in cerebral
ischaemia
Because fibrates, used as lipid-lowering agents, contribute to
secondary prevention of stroke, it has been supposed that
these PPARα activators could also preventively protect the
brain against noxious biological reactions induced by cerebral
C
2006 Biochemical Society
1344 Biochemical Society Transactions (2006) Volume 34, part 6
ischaemia, such as oxidative stress and inflammation. It has
been demonstrated that a 14-day preventive treatment with
fenofibrate reduced susceptibility to stroke in apolipoprotein
E-deficient mice as well as decreased cerebral infarct volume
in wild-type mice [18]. In another study, it was confirmed that
two different PPARα agonists, fenofibrate and Wy-14643,
provided similar brain protection when administered res-
pectively 3 or 7 days before induction of cerebral ischaemia
[19]. More recently, it has been demonstrated that PPARα
agonists could also induce an acute neuroprotection when
administered just before cerebral ischaemia or during the
reperfusion period [20,21].
Administration of the PPARγ agonists troglitazone or
pioglitazone 24 or 72 h before and at the time of cerebral
infarction dramatically reduced infarction volume and
improved neurological function following transient middle
cerebral artery occlusion in rats [22,23]. This effect is exerted
in a dose-dependent manner. This neuroprotection has been
reproduced by an intracerebroventricular administration of
pioglitazone, proving that it is the activation of intracerebral
PPARγ that confers neuroprotection and neurological
improvement following ischaemic injury [24]. Moreover,
a non-thiazolidinedione PPARγ agonist (L-796449) also
had a neuroprotective effect in experimental stroke and was
also found to activate 15-deoxy-
12,14
-prostaglandin J
2
by
adenoviral transfer of COX-2 [25,26]. In a first study, PPARγ
agonists had no effect in a permanent model of cerebral
ischaemia, suggesting that mechanisms of action could take
place during reperfusion [23], while recent results give the
opposite effect [25].
Mechanisms of PPAR-induced neuroprotection
Cerebral mechanisms
The neuroprotection observed after treatment with PPAR
agonists is related to several mechanisms including both
oxidative stress modulation and anti-inflammatory effect.
PPARα agonist-induced neuroprotective effect is associated
with a decrease in cerebral oxidative stress depending on
the increase in activity of numerous antioxidant enzymes,
in particular Cu/Zn superoxide dismutase and glutathione
peroxidase [18]. This modulation of antioxidant enzymes
is responsible for a decrease in ischaemia-induced reactive
oxygen species production and lipid peroxidation [21,27].
This effect on oxidative stress could be related to a direct
effect on antioxidant enzymes expression, because PPREs
(PPAR-response elements) have been found in the gene of
Cu/Zn superoxide dismutase [5].
The neuroprotective effects of PPAR agonists are also
related to inhibition of ischaemia-induced inflammatory
markers (interleukin-1β, COX-2 and inducible nitric oxide
synthase) [21,27]. The different PPAR isoforms do not modu-
late the inflammatory pathways involved in neuroprotection
in a similar manner. For instance, ischaemia-induced COX-2
overexpression is prevented by PPARγ agonists but not by
PPARα agonists [21,22,27]. There is a link between PPAR-
induced modulation of oxidative stress and inflammation,
since prevention of COX-2 induction results from oxidative
stress inhibition [28]. The cellular target of these anti-
inflammatory effects is probably microglial cells, since
PPARγ agonists, such pioglitazone, are able to decrease mi-
croglial activation when administered intracerebrally [24,29].
The key target of this anti-inflammatory effect is NF-κB,
which plays a crucial role in neuronal death [30]. PPARγ and
PPARα activation is responsible for inhibition of the NF-κB
p65 monomer as well as induction of IκBα (inhibitory κB)
[25,31]. The role of suppression of activation of p38 mitogen-
activated protein kinase has also been demonstrated recently
[21,27].
Beyond this direct effect on ischaemia-induced deleterious
pathways explaining neuroprotection, the challenge will be
to demonstrate that a part of the neurological improvement
induced by PPAR activators could be the result of
neurorepair, since P PARγ s are also involved in the regulation
of neural stem cell proliferation and differentiation [32].
Vascular mechanisms
Because PPARs are mainly expressed in cerebral vascular
wall, in particular in endothelium, it has been supposed that
vascular mechanisms could be involved in neuroprotection.
Thus preventive neuroprotection by PPARα is associated
with an improvement in middle cerebral artery sensitivity
to endothelium-dependent relaxation unrelated to an in-
crease in endothelial nitric oxide synthase expression [18].
More recently, it has been demonstrated that preventive
or acute PPARα agonist-induced neuroprotection paral-
leled the prevention of ischaemia-induced endothelial dys-
function [20]. This vascular effect could be related to: (i) the
prevention of ischaemia-induced vascular expression of
adhesion molecules; (ii) the antioxidant effect of PPAR
activation; and (iii) the inhibition of ischaemia-induced
metalloproteinase expression [18,25]. In addition, PPAR
could also be involved in endothelial regeneration as has been
demonstrated in other arterial areas [33].
PPAR and neuroprotection: beyond
cerebral ischaemia
Other acute cerebral injuries such as traumatic brain injury
or chronic neurological diseases such as neurodegenerative
diseases or multiple sclerosis also need pleiotropic neuro-
protective drugs, explaining why PPAR activators have also
been tested in experimental models mimicking these different
disorders.
Traumatic brain and spinal cord injury
Because many mechanisms that are involved in cerebral
ischaemia are also involved in traumatic nervous tissue injury,
the effect of PPARα activation has been tested in models of
traumatic spinal cord and brain injury, and a neuroprotective
effect has been observed with some similar mechanisms to
those in cerebral ischaemia [34,35].
Alzheimer’s disease
PPARγ and PPARα agonists have been tested in models of
Alzheimer’s disease. The classical histopathological hallmarks
C
2006 Biochemical Society
International Symposium on Neurodegeneration and Neuroprotection 1345
of Alzheimer’s disease include extracellular Aβ (amyloid
β-peptide) deposition in neuritic plaques and intracellular
deposits of hyperphosphorylated tau protein, causing form-
ation of neurofibrillary tangles and finally neuronal death,
responsible for progressive memory loss and decline of
cognitive functions. While it has been demonstrated that
aPPARα agonist inhibited Aβ-stimulated expression of
tumour necrosis factor α and interleukin-6 reporter genes in a
dose-dependent manner, but failed to inhibit Aβ-stimulated
elaboration of neurotoxic factors [36], some recent experi-
mental data suggest that fenofibrate could raise Aβ-(1–42)
production [37], suggesting that PPARα remains a contro-
versial target in Alzheimer’s disease. PPARγ agonists were
also shown to inhibit the β-amyloid-stimulated expression
of inflammatory cytokines and COX-2 [38]. In addition to
inhibition of Aβ-induced inflammation, PPARγ could also
induce clearance of the β-amyloid peptide [39]. In addition to
in vitro data, recent in vivo data also indicate a beneficial effect
of PPARγ activation, since an acute 7-day oral treatment
with the PPARγ agonist pioglitazone resulted in a reduction
in glial activation as well as a reduction in the number of
Aβ-positive plaque areas in the hippocampus and cortex
of a murine transgenic model of the amyloid pathology of
Alzheimer’s disease [40].
Parkinson’s disease
PPAR agonists have also been assessed in a model of
Parkinson’s disease. Parkinson’s disease is characterized by
a progressive loss of dopaminergic neurons in the substantia
nigra, which is experimentally mimicked by systemic
administration of the neurotoxin MPTP (1-methyl-4-phenyl-
1,2,3,6-tetrahydropyridine). Oral administration of the
PPARγ agonist pioglitazone attenuated the MPTP-induced
glial activation and prevented dopaminergic cell loss in the
substantia nigra pars compacta. Pioglitazone also prevented
MPTP-induced expression of inducible nitric oxide synthase
[41]. This protective effect of pioglitazone is also associated
with an increase in inhibitory protein-κBα expression and to
inhibition of translocation of the NF-κB subunit p65 to the
nucleus in dopaminergic neurons, glial cells and astrocytes
[42]. Preliminary results demonstrate that PPARα activation
prevents death of dopaminergic neurons of substantia nigra
pars compacta in the MPTP model of Parkinson’s disease [43].
Multiple sclerosis
Microglial activation and inflammation are the key to the
pathophysiology of multiple s clerosis, explaining why PPAR
agonists have been tested in this disease, in particular in
the model of experimental autoimmune EAE (encephalo-
myelitis), which is characterized by CNS inflammation and
demyelination, together with remittent paralysis [16]. Oral
administration of gemfibrozil and fenofibrate, two PPARα
agonists, also inhibits clinical signs of EAE by mechanisms
involving secretion of interferon-γ and interleukin-4 [44].
Oral administration of the PPARγ agonist pioglitazone
reduces the motor symptoms’ severity in monophasic EAE,
without delaying the disease onset. In a relapsing model of
EAE, pioglitazone reduces the severity of relapses and overall
mortality without affecting t he onset and severity of the
initial disease attack [45]. The mechanisms of action of PPAR
agonists in EAE are complex, involving regulation of the
inflammatory pathway and also modulation of the maturation
and differentiation of oligodendrocytes [16].
Conclusion
The hypothesis that the pleiotropic effects of PPAR
agonist could decrease neuronal death is supported by
much experimental data showing that PPAR agonists exert
neuroprotective effects in models of cerebral ischaemia,
neurodegenerative diseases and multiple sclerosis, with some
clinical data confirming these experimental results. These res-
ults have been essentially obtained with PPARγ and PPARα
activators, while the PPARβ/δ pathway remains largely
unexplored despite interest in the target. Development of new
and more potent PPAR activators as well as combined action
of the different isoforms of PPAR are also future prospects
in terms of neuroprotection and also in terms of neurorepair.
References
1 Cheng, Y.D., Al-Khoury, L. and Zivin, J.A. (2004) NeuroRx. 1, 36–45
2 Liebeskind, D.S. and Kasner, S.E. (2001) CNS Drugs. 15, 165–174
3 Lees, K.R., Zivin, J.A. and Ashwood, T. (2006) New Engl. J. Med. 354,
588–600
4 Berger, J.P., Akiyama, T.E. and Meinke, P.T. (2005) Trends Pharmacol. Sci.
26, 244–251
5 Moraes, L., Piqueras, L. and Bishop-Bailey, D. (2006) Pharmacol. Ther.
110, 371–385
6 Blanquart, C., Barbier, O., Fruchart, J.C., Staels, B. and Glineur, C. (2003)
J. Steroid Biochem. Mol. Biol. 85, 267–273
7 Bishop-Bailey, D. (2000) Br. J. Pharmacol. 129, 823–834
8 Buchan, K.W. and Hassall, D.G. (2000) Med. Res. Rev. 20, 350–366
9 Luquet, S., Lopez-Soriano, J., Holst, D., Gaudel, C., Jehl-Pietri, C.,
Fredenrich, A. and Grimaldi, P.A. (2004) Biochimie 86, 833–837
10 Cullingford, T.E., Bhakoo, K., Peuchen, S., Dolphin, C.T., Patel, R. and
Clark, J.B. (1998) J. Neurochem. 70, 1366–1375
11 Moreno, S., Farioli-Vecchioli, S. and Cer `u, M.P. (2004) Neuroscience 123,
131–145
12 Andersen, J.K. (2004) Nat. Med. 10, S18–S25
13 Feinstein, D.L. (2003) Diabetes Technol. Ther. 5, 67–73
14 Kim, E.J., Kwon, K.J., Park, J.Y., Lee, S.H., Moon, C.H. and Baik, E.J. (2002)
Brain Res, 941,110
15 Zhao, X., Ou, Z., Gotta, J.C., Waxham, N. and Aronowski, J. (2006)
Brain Res. 1073–1074, 460–469
16 Kielian, T. and Drew, P.D. (2003) J. Neurosci. Res. 71, 315–325
17 Uryu, S., Harada, J., Hisamoto, M. and Oda, T. (2002) Brain Res. 924,
229–236
18 Deplanque, D., Gel ´e, P., P ´etrault, O., Six, I., Furman, C., Bouly, M., Nion, S.,
Dupuis, B., Leys, D., Fruchart, J.C. et al. (2003) J. Neurosci. 23, 6264–6271
19 Inoue, H., Jiang, X.-F., Katayama, T., Osada, S., Umesono, K. and
Namura, S. (2003) Neurosci. Lett. 352, 203–206
20 Ouk, T., Petrault, O., Gautier, S., Gel ´e, P., Laprais, M., Duriez, P.,
Bastide, M. and Bordet, R. (2005) J. Cereb. Blood Flow Metab. 25,
(Suppl. 1), S56
21 Collino, M., Aragno, M., Mastrocola, R., Benetti, E., Gallicchio, M.,
Dianzani, C., Danni, O., Thiemermann, C. and Fantozzi, R. (2006)
Free Radical Biol. Metab. 41, 579–589
22 Sundararajan, S., Gamboa, J.L., Victor, N.A., Wanderi, E.W., Lust, W.D. and
Landreth, G.E. (2005) Neuroscience 130, 685–696
23 Shimazu, T., Inoue, I., Araki, N., Asano, Y., Sawada, M., Furuya, D.,
Nagoya, H. and Greenberg, J.H. (2005) Stroke 36, 353–359
24 Zhao, Y., Patzer, A., Gohike, P., Herdegen, T. and Culman, J. (2005)
Eur. J. Neurosci. 22, 278–282
C
2006 Biochemical Society
1346 Biochemical Society Transactions (2006) Volume 34, part 6
25 Pereira, M.P., Hurtado, O., Cardenas, A., Alonso-Escolano, D., Bosc ´a, L.,
Vivancos, J., Nombela, F., Leza, J.C., Lorenzo, P., Lizasoain, I. and Moro,
M.A. (2005) J. Neuropathol. Exp. Neurol. 64, 797–805
26 Lin, T.N., Cheung, W.M., Wu, J.S., Chen, J.J., Lin, H., Chen, J.J., Liou, J.Y.,
Shyue, S.K. and Wu, K.K. (2006) Arterioscl. Thromb. Vasc. Biol. 26,
481–487
27 Collino, M., Aragno, M., Mastrocola, R., Benetti, E., Gallicchio, M., Rosa,
A.C., Dianzani, C., Danni, O., Thiemermann, C. and Fantozzi, R. (2006)
Eur. J. Pharmacol. 430, 70–80
28 Zhao, Y., Patzer, A., Gohike, P., Herdegen, T. and Culman, J. (2006)
FASEB J. 20, 1162–1175
29 Bernardo, A., Gasparini, L., Ongini, E. and Minghetti, L. (2006)
Neurobiol. Dis. 22, 25–32
30 Hu, X., Nesic-Taylor, O., Qiu, J., Rea, H.C., Fabian, R., Rassin, D.K. and
Perez-Polo, J.R. (2005) J. Neurochem. 93, 26–37
31 Zambon, A., Gervois, P., Pauletto, P., Fruchart, J.-C. and Staels, B. (2006)
Arterioscl. Thromb. Vasc. Biol. 26, 977–986
32 Wada, K., Nakajima, A., Katayama, K., Kudo, C., Shibuya, A., Kubota, N.,
Terauchi, M., Miyoshi, H., Kamisaki, Y., Mayumi, T. et al. (2006)
J. Biol. Chem. 281, 12673–12681
33 Tanaka, T., Fukunaga, Y., Itoh, H., Doi, K., Yamashita, J., Chun, T-H.,
Inoue, M., Masatsugu, K., Saito, T., Sawada, N. et al. (2005)
Eur. J. Pharmacol. 508, 255–265
34 Genovese, T., Mazzon, E., Di Paola, R., Cannavo, G., Muia, C., Bramanti, P.
and Cuzzocrea, S. (2005) Exp. Neurol. 194, 267–278
35 Besson, V.C., Chen, X.P., Plotkine, M. and Marchand-Verrechia, C. (2005)
Neurosci. Lett. 388,712
36 Combs, C.K., Bates, P., Karlo, J.C. and Landreth, G.E. (2001)
Neurochem. Int. 39, 449–457
37 Kukar, T., Murphy, M.P., Eriksen, J.L., Sagi, S.A., Weggen, S., Smith, T.E.,
Ladd, T., Khan, M.A., Kache, R., Beard, J. et al. (2005) Nat. Med. 11,
545–550
38 Combs, C.K., Johnson, D.E., Karlo, J.C., Cannady, S.B. and Landreth, G.E.
(2000) J. Neurosci. 20, 558–567
39 Camacho, I.E., Serneels, L., Spittaels, K., Merchiers, P., Dominguez, D. and
De Strooper, B. (2004) J. Neurosci. 24, 10908–10917
40 Heneka, M.T., Sastre, M., Dumitrescu-Ozimek, L., Hanke, A.,
Dewachter, I., Kuiperi, C., O’Banion, K., Klockgether, T.,
Van Leuven, F. and Landreth, G.E. (2005) Brain 128, 1442–1453
41 Breidert, T., Callebert, J., Heneka, M.T., Landreth, G., Launay, J.M. and
Hirsch, E.C. (2002) J. Neurochem. 82, 615–624
42 Dehmer, T., Heneka, M.T., Sastre, M., Dichgans, J. and Schulz, J.B. (2004)
J. Neurochem. 88, 494–501
43 Kreisler, A., Gele, P., Destee, A. and Bordet, R. (2003)
Fundam. Clin. Pharmacol. 17, 262
44 Lovett-Racke, A.E., Hussain, R.Z., Northrop, S., Choy, J., Rocchini, A.,
Matthes, L., Chavis, J.A., Diab, A., Drew, P.D. and Racke, M.K. (2004)
J. Immunol. 172, 5790–5798
45 Feinstein, D.L., Galea, E., Gavrilyuk, V., Brosnan, C.F., Whitacre, C.C.,
Dumitrescu-Ozimek, L., Landreth, G.E., Pershadsingh, H.A., Weinberg, G.
and Heneka, M.T. (2002) Ann. Neurol. 51, 694–702
Received 11 August 2006
C
2006 Biochemical Society
... ППАРс реэулате тарэет эенес ехпрессион [9]. Мост реъент дисъовериес портрай ППАРс ас промисинэ пщармаъолоэиъал тарэетс фор тще треатмент оф аъуте исъщемиъ строке, тщанкс то тщеир абилитй то симултанеоуслй интерфере wитщ северал меъщанисмс тщат ундерлие тще патщопщйсиолоэй оф браин исъщемиа, тщус леадинэ то ан интерестинэ протеътиве стратеэй то ъоунтераът тще мултипле делетериоус еффеътс оф исъщемиъ анд пост-исъщемиъ инжуриес [11]. Тщрее тйпес оф ППАРс щаве беен идентифиед -алпща, делта/бета анд эамма. ...
Article
Stroke is a leading cause of death and disability among adult population. Many pathological events including inflammation, oxidative stress, and apoptosis contribute to the secondary neuronal death after stroke. The goal of this review is to discuss the therapeutic potential and putative mechanisms of neuroprotective properties of thiazolidinediones (peroxisome proliferator-activated receptors-? agonists) at ischemic stroke. Thiazolidinediones have insulin-sensitizing and other additional pleiotropic properties. Ischemic neuroprotection afforded by thiazolidinediones has been involved anti-inflammatory, anti-oxidant, anti-apoptotic properties, as well as effects on endothelial function and repair. These novel actions of thiazolidinediones could offer some protection against the potentially enhanced damage of brain ischemia in patients with abdominal obesity and insulin resistance and may open new exciting lines of investigation on stroke treatment.
... PPAR γ, a nuclear receptor located in certain parts of the brain, plays a vital role in glucose and lipid metabolism [79,80]. PPARγ also increases neuronal inflammation and injury [81]. Inhibiting this receptor decreases Aβ aggregates and neuroinflammatory mediator expressions [82,83]. ...
Article
Alzheimer’s disease (AD) is a very common neurodegenerative disorder associated with memory loss and a progressive decline in cognitive activity. The two major pathophysiological factors responsible for AD are amyloid plaques (comprising amyloid-beta aggregates) and neurofibrillary tangles (consisting of hyperphosphorylated tau protein). Polyphenols, a class of naturally occurring compounds, are immensely beneficial for the treatment or management of various disorders and illnesses. Naturally occurring sources of polyphenols include plants and plant-based foods, such as fruits, herbs, tea, vegetables, coffee, red wine, and dark chocolate. Polyphenols have unique properties, such as being the major source of anti-oxidants and possessing anti-aging and anti-cancerous properties. Currently, dietary polyphenols have become a potential therapeutic approach for the management of AD, depending on various research findings. Dietary polyphenols can be an effective strategy to tackle multifactorial events that occur with AD. For instance, naturally occurring polyphenols have been reported to exhibit neuroprotection by modulating the Aβ biogenesis pathway in AD. Many nanoformulations have been established to enhance the bioavailability of polyphenols, with nanonization being the most promising. This review comprehensively provides mechanistic insights into the neuroprotective potential of dietary polyphenols in treating AD. It also reviews the usability of dietary polyphenol as nanoformulation for AD treatment.
... The increased intracellular concentration of glutamate can lead to the rearrangement of oxidative metabolism and accelerate the plasticity of macrophages toward immunosuppressive phenotypes. Microglia and astrocytes are critical in the development of neuroinflammatory diseases (Fakhoury, 2018;Palpagama et al., 2019), suggesting that regulation of PPAR-γ and EAAT2 can play an important role in attenuating neurodegenerative processes in these cells (Bordet et al., 2016). ...
Article
Full-text available
Purpose During the course of demyelinating inflammatory diseases, myelin-derived proteins, including myelin basic protein(MBP), are secreted into extracellular space. MBP shows extensive post-translational modifications, including deimination/citrullination. Deiminated MBP is structurally less ordered, susceptible to proteolytic attack, and more immunogenic than unmodified MBP. This study investigated the effect of the deiminated/citrullinated isomer of MBP(C8) and the unmodified isomer of MBP(C1) on cultured primary astrocytes. Methods MBP charge isomers were isolated/purified from bovine brain. Primary astrocyte cultures were prepared from the 2-day-old Wistar rats. For evaluation of glutamate release/uptake a Fluorimetric glutamate assay was used. Expression of peroxisome proliferator-activated receptor-gamma(PPAR-γ), excitatory amino acid transporter 2(EAAT2), the inhibitor of the nuclear factor kappa-B(ikB) and high mobility group-B1(HMGB1) protein were assayed by Western blot analysis. IL-17A expression was determined in cell medium by ELISA. Results We found that MBP(C8) and MBP(C1) acted differently on the uptake/release of glutamate in astrocytes: C1 increased glutamate uptake and did not change its release, whereas C8 decreased glutamate release but did not change its uptake. Both isomers increased the expression of PPAR-γ and EAAT2 to the same degree. Western blots of cell lysates revealed decreased expression of ikB and increased expression of HMGB1 proteins after treatment of astrocytes by C8. Moreover, C8-treated cells released more nitric oxide and proinflammatory IL-17A than C1-treated cells. Conclusions These data suggest that the most immunogenic deiminated isomer C8, in parallel to the decreases in glutamate release, elicits an inflammatory response and enhances the secretion of proinflammatory molecules via activation of nuclear factor kappa B(NF-kB). Summary statement The most modified-citrullinated myelin basic protein charge isomer decreases glutamate release, elicits an inflammatory response and enhances the secretion of proinflammatory molecules via activation of nuclear factor kappa B in astrocytes.
... PPARα is a transcription factor that belongs to the nuclear receptor superfamily (37). Upon activation, PPARα heterodimerizes with the Retinoid X Receptor (RXR) and binds to PPAR Response Elements (PPREs) in the promoter regions of target genes including PPARα itself and those involved in many processes such as energy metabolism, oxidative stress, inflammation, circadian rhythm, immune response, mitochondrial genesis, and cell differentiation (12,(38)(39)(40)(41)(42)(43). Two independent, prospective clinical studies reported robust therapeutic effects of PPARα agonist fenofibrate on diabetic retinopathy in type 2 diabetic patients (44). ...
Article
Full-text available
Diabetes can result in impaired corneal wound healing. Mitochondrial dysfunction plays an important role in diabetic complications. However, the regulation of mitochondria function in the diabetic cornea and its impacts on wound healing remain elusive. The present study aimed to explore the molecular basis for the disturbed mitochondrial metabolism and subsequent wound healing impairment in the diabetic cornea. Seahorse analysis showed that mitochondrial oxidative phosphorylation is a major source of ATP production in human corneal epithelial cells. Live corneal biopsy punches from type 1 and type 2 diabetic mouse models showed impaired mitochondrial functions, correlating with impaired corneal wound healing, compared to nondiabetic controls. To approach the molecular basis for the impaired mitochondrial function, we found that Peroxisome Proliferator-Activated Receptor-α (PPARα) expression was downregulated in diabetic human corneas. Even without diabetes, global PPARα knockout mice and corneal epithelium-specific PPARα conditional knockout mice showed disturbed mitochondrial function and delayed wound healing in the cornea, similar to that in diabetic corneas. In contrast, fenofibrate, a PPARα agonist, ameliorated mitochondrial dysfunction and enhanced wound healing in the corneas of diabetic mice. Similarly, corneal epithelium-specific PPARα transgenic overexpression improved mitochondrial function and enhanced wound healing in the cornea. Furthermore, PPARα agonist ameliorated the mitochondrial dysfunction in primary human corneal epithelial cells exposed to diabetic stressors, which was impeded by siRNA knockdown of PPARα, suggesting a PPARα-dependent mechanism. These findings suggest that downregulation of PPARα plays an important role in the impaired mitochondrial function in the corneal epithelium and delayed corneal wound healing in diabetes.
... While exenatide has shown promising data in the above study, another diabetes drug, pioglitazone, did not modify progression in early PD in a different Phase 2 trial published in 2015 [282]. The mechanism of action of pioglitazone differs from that of exenatide in that the former is an agonist of peroxisome proliferator-activated receptors (PPARs) PPARα and PPARγ [283]. Currently, it is not clear whether the difference in efficacies against PD between pioglitazone and exenatide is due to the two drugs targeting different aspects of insulin signaling or other less obvious reasons. ...
Article
Full-text available
α-synuclein is a pleiotropic protein underlying a group of progressive neurodegenerative diseases, including Parkinson’s disease and dementia with Lewy bodies. Together, these are known as synucleinopathies. Like all neurological diseases, understanding of disease mechanisms is hampered by the lack of access to biopsy tissues, precluding a real-time view of disease progression in the human body. This has driven researchers to devise various experimental models ranging from yeast to flies to human brain organoids, aiming to recapitulate aspects of synucleinopathies. Studies of these models have uncovered numerous genetic modifiers of α-synuclein, most of which are evolutionarily conserved. This review discusses what we have learned about disease mechanisms from these modifiers, and ways in which the study of modifiers have supported ongoing efforts to engineer disease-modifying interventions for synucleinopathies.
Article
Full-text available
Citation: Ibáñez, C.; Acuña, T.; Quintanilla, M.E.; Pérez-Reytor, D.; Morales, P.; Karahanian, E. Fenofibrate Decreases Ethanol-Induced Neuroinflammation and Oxidative Stress and Reduces Alcohol Relapse in Rats by a PPAR-α-Dependent Mechanism. Abstract: High ethanol consumption triggers neuroinflammation, implicated in sustaining chronic alcohol use. This inflammation boosts glutamate, prompting dopamine release in reward centers, driving prolonged drinking and relapse. Fibrate drugs, activating peroxisome proliferator-activated receptor alpha (PPAR-α), counteract neuroinflammation in other contexts, prompting investigation into their impact on ethanol-induced inflammation. Here, we studied, in UChB drinker rats, whether the administration of fenofibrate in the withdrawal stage after chronic ethanol consumption reduces voluntary intake when alcohol is offered again to the animals (relapse-type drinking). Furthermore, we determined if fenofibrate was able to decrease ethanol-induced neuroinflammation and oxidative stress in the brain. Animals treated with fenofibrate decreased alcohol consumption by 80% during post-abstinence relapse. Furthermore, fenofibrate decreased the expression of the proinflammatory cytokines tumor necrosis factor-alpha (TNF-α) and interleukins IL-1β and IL-6, and of an oxidative stress-induced gene (heme oxygenase-1), in the hippocampus, nucleus accumbens, and prefrontal cortex. Animals treated with fenofibrate showed an increase M2-type microglia (with anti-inflammatory proprieties) and a decrease in phagocytic microglia in the hippocampus. A PPAR-α antagonist (GW6471) abrogated the effects of fenofibrate, indicating that they are dependent on PPAR-α activation. These findings highlight the potential of fenofibrate, an FDA-approved dyslipidemia medication, as a supplementary approach to alleviating relapse severity in individuals with alcohol use disorder (AUD) during withdrawal.
Article
This review highlights the potential role of cyclooxygenase-2 enzyme (COX-2) in the pathogenesis of Alzheimer's disease (AD) and the potential therapeutic use of non-steroidal anti-inflammatory drugs (NSAIDs) in the management of AD. In addition to COX-2 enzymes role in inflammation, the formation of amyloid plaques and neurofibrillary tangles in the brain, the review emphasizes that COXs-2 have a crucial role in normal synaptic activity and plasticity, and have a relationship with acetylcholine, tau protein, and beta-amyloid (Aβ) which are the main causes of Alzheimer's disease. Furthermore, the review points out that COX-2 enzymes have a relationship with kinase enzymes, including Cyclin Dependent Kinase 5 (CDK5) and Glycogen Synthase Kinase 3β (GSK3β), which are known to play a role in tau phosphorylation and are strongly associated with Alzheimer's disease. Therefore, the use of drugs like NSAIDs may be a hopeful approach for managing AD. However, results from studies examining the effectiveness of NSAIDs in treating AD have been mixed and further research is needed to fully understand the mechanisms by which COX-2 and NSAIDs may be involved in the development and progression of AD and to identify new therapeutic strategies.
Article
Full-text available
Astrocyte inflammation activation is an important cause that hinders the recovery of motor function after cerebral ischemia. However, its molecular mechanism has not yet been clearly clarified. The peroxisome proliferator-activated receptor α (PPARα) is a ligand-activated nuclear transcriptional factor. This study aims to further clarify the role of PPARα in astrocyte inflammation activation after cerebral ischemia and to explore the underlying mechanism. Astrocyte activation was induced in an in vivo model by transient middle cerebral artery occlusion (tMCAO) in mice. The in vitro model was induced by an oxygen-glucose deprivation/reoxygenation (OGD/R) in a primary culture of mouse astrocyte. PPARα-deficient mice were used to observe the effects of PPARα on astrocyte activation and autophagic flux. Our results showed that PPARα was mainly expressed in activated astrocytes during the chronic phase of brain ischemia and PPARα dysfunction promoted astrocyte inflammatory activation. After cerebral ischemia, the expressions of LC3-II/I and p62 both increased. Autophagic vesicle accumulation was observed by electron microscopy in astrocytes, and the block of autophagic flux was indicated by an mRFP-GFP-LC3 adenovirus infection assay. A PPARα deficit aggravated the autophagic flux block, while PPARα activation preserved the lysosome function and restored autophagic flux in astrocytes after OGD/R. The autophagic flux blocker bafilomycin A1 and chloroquine antagonized the effect of the PPARα agonist on astrocyte activation inhibition. This study identifies a potentially novel function of PPARα in astrocyte autophagic flux and suggests a therapeutic target for the prevention and treatment of chronic brain ischemic injury.
Chapter
Sensory neurons are critical targets in diabetes mellitus (DM). Diabetic polyneuropathy (DPN) can be considered a unique form of sensory predominant neurodegeneration that renders loss of distal axon terminals, especially those in the skin, with relative preservation of their cell body (perikarya). Patients experience loss of sensation (numbness), gait instability with falls, unrecognized injury of insensate limbs with skin ulceration, and neuropathic pain. Sensory neurons reside in paraspinal dorsal root (and trigeminal) ganglia (DRGs) possessing unique microvascular and barrier properties that lead to greater vulnerability from DM. The molecular responses of sensory neurons differ from those of axotomized peripheral neurons. In DM the changes emphasize downregulation of key structural proteins, shifts in ion channel expression, and attenuated growth proteins all indicative of chronic neurotoxic stress. Changes in several differentially expressed mRNAs and miRNAs of DRG neurons in DPN, such as CWC22 and mmu-Let-7i, may contribute to sensory dysfunction. Finally, molecular strategies emphasizing regenerative impacts, including topical approaches, have the capacity to reverse features of DPN including loss of skin innervation. These have included local insulin (intrathecal, intranasal, near nerve, intrahindpaw) given in doses that do not alter hyperglycemia, GLP-1 agonists, PTEN (phosphatase and tensin homolog deleted on chromosome 10) inhibition or knockdown, and muscarinic antagonists. Several additional and novel strategies are emerging that may influence axonal degeneration of distal sensory terminals or axon regeneration specifically. Despite a limited clinical trial track record over several decades, new mechanistic insights for translation in DPN offer hope for better trial results.KeywordsDiabetic polyneuropathySensory neuronDorsal root gangliaAxon regenerationSensory neurodegenerationAxon terminalSensory perikaryaDifferential RNA expression
Article
Full-text available
Alzheimer's disease (AD) is characterized by the extracellular deposition of beta-amyloid fibrils within the brain and the subsequent association and phenotypic activation of microglial cells associated with the amyloid plaque. The activated microglia mount a complex local proinflammatory response with the secretion of a diverse range of inflammatory products. Nonsteroidal anti-inflammatory drugs (NSAIDs) are efficacious in reducing the incidence and risk of AD and significantly delaying disease progression. A recently appreciated target of NSAIDs is the ligand-activated nuclear receptor peroxisome proliferator-activated receptor gamma (PPARgamma). PPARgamma is a DNA-binding transcription factor whose transcriptional regulatory actions are activated after agonist binding. We report that NSAIDs, drugs of the thiazolidinedione class, and the natural ligand prostaglandin J2 act as agonists for PPARgamma and inhibit the beta-amyloid-stimulated secretion of proinflammatory products by microglia and monocytes responsible for neurotoxicity and astrocyte activation. The activation of PPARgamma also arrested the differentiation of monocytes into activated macrophages. PPARgamma agonists were shown to inhibit the beta-amyloid-stimulated expression of the cytokine genes interleukin-6 and tumor necrosis factor alpha. Furthermore, PPARgamma agonists inhibited the expression of cyclooxygenase-2. These data provide direct evidence that PPARgamma plays a critical role in regulating the inflammatory responses of microglia and monocytes to beta-amyloid. We argue that the efficacy of NSAIDs in the treatment of AD may be a consequence of their actions on PPARgamma rather than on their canonical targets the cyclooxygenases. Importantly, the efficacy of these agents in inhibiting a broad range of inflammatory responses suggests PPARgamma agonists may provide a novel therapeutic approach to AD.
Article
Successful management of cardiovascular (CV) disease and associated metabolic syndromes, such as diabetes, is a major challenge to the clinician. Reducing CV risk factors, such as abnormal lipid profiles, insulin resistance or hypertension is the foundation of such therapy. A relatively new class of therapeutic agent, activators of peroxisome proliferator-activated receptors (PPAR), is poised to make a major impact with regard to several areas of risk factor management. However, there is growing evidence that PPAR agonists may also influence the CV system directly by modulating vessel wall function. These observations suggest that additional benefit, in the treatment of CV disease, may derive not only from the ability of agents to modify risk factors but also to influence directly the cellular mechanisms of disease within the vessel wall. A precedent for this dual action comes from examination of the effects of inhibitors of HMG CoA reductase (statins), where risk factor modulation is accompanied by direct actions on the vessel wall. In this review, we summarize the evidence suggesting that PPAR agonists may directly modulate vessel wall function, and that these may parallel those effects reported recently for the statins, (C) 2000 John Wiley & Sons, Inc.
Article
Peroxisome proliferator-activated receptors (PPARs) are ligand-dependent transcription factors which belong to the nuclear receptor family. We examined whether PPARα agonists and resveratrol, a polyphenol contained in grapes, protect the brain against ischemia. To investigate whether resveratrol activates PPARs, we performed a cell-based transfection activity assay using luciferase reporter plasmid. PPARα and PPARγ were activated by resveratrol in primary cortical cultures and vascular endothelial cells. Resveratrol (20 mg/kg, 3 days) reduced infarct volume by 36% at 24 h after middle cerebral artery occlusion in wild-type mice. The PPARα agonists fenofibrate (30 mg/kg, 3 days) and Wy-14643 (30 mg/kg, days) exerted similar brain protection. However, resveratrol and fenofibrate failed to protect the brain in PPARα knockout mice. The data indicate that PPARα agonists protect the brain through PPARα.
Article
Inflammation has been implicated in the pathogenesis of Parkinson's disease (PD). In the chronic 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP) model of PD, inducible NO synthase (iNOS) derived nitric oxide (NO) is an important mediator of dopaminergic cell death. Ligands of the peroxisome proliferator-activated receptor (PPAR) exert anti-inflammatory effects. We here investigated whether pioglitazone, a PPARγ agonist, protected mice from MPTP-induced dopaminergic cell loss, glial activation, and loss of catecholamines in the striatum. As shown by western blot, PPARγ was expressed in the striatum and the substantia nigra of vehicle- and MPTP-treated mice. Oral administration of 20 mg/(kg day) of pioglitazone protected tyrosine hydroxylase (TH)-positive substantia nigra neurons from death induced by 5 × 30 mg/kg MPTP. However, the decrease of dopamine in the striatum was only partially prevented. In mice treated with pioglitazone, there were a reduced activation of microglia, reduced induction of iNOS-positive cells and less glial fibrillary acidic protein positive cells in both striatum and substantia nigra pars compacta. In addition, treatment with pioglitazone almost completely blocked staining of TH-positive neurons for nitrotyrosine, a marker of NO-mediated cell damage. Because an increase in inhibitory protein-κ-Bα (IκBα) expression and inhibition of translocation of the nuclear factor kappaB (NFκB) subunit p65 to the nucleus in dopaminergic neurons, glial cells and astrocytes correlated with the protective effects of pioglitazone, our results suggest that pioglitazone sequentially acts through PPARγ activation, IκBα induction, block of NFκB activation, iNOS induction and NO-mediated toxicity. In conclusion, treatment with pioglitazone may offer a treatment opportunity in PD to slow the progression of disease that is mediated by inflammation.
Article
We report the isolation, by RT-PCR, of partial cDNAs encoding the rat peroxisome proliferator-activated receptor (PPAR) isoforms PPAR, PPARβ, and PPAR and the rat retinoid X receptor (RXR) isoforms RXR, RXRβ, and RXR. These cDNAs were used to generate antisense RNA probes to permit analysis, by the highly sensitive and discriminatory RNase protection assay, of the corresponding mRNAs in rat brain regions during development. PPAR, PPARβ, RXR, and RXRβ mRNAs are ubiquitously present in different brain regions during development, PPAR mRNA is essentially undetectable, and RXR mRNA is principally localised to cortex. We demonstrate, for the first time, the presence of PPAR and RXR mRNAs in primary cultures of neonatal meningeal fibroblasts, cerebellar granule neurons (CGNs), and cortical and cerebellar astrocytes and in primary cultures of adult cortical astrocytes. PPAR, PPARβ, RXR, and RXRβ mRNAs are present in all cell types, albeit that PPAR and RXR mRNAs are at levels near the limit of detection in CGNs. PPAR mRNA is expressed at low levels in most cell types but is present at levels similar to those of PPAR mRNA in adult astrocytes. RXR mRNA is present either at low levels, or below the level of detection of the assay, for all cell types studied.
Article
Peroxisome proliferator-activated receptor (PPAR)s are a family of three nuclear hormone receptors, PPARα, -δ, and -γ, which are members of the steriod receptor superfamily. The first member of the family (PPARα) was originally discovered as the mediator by which a number of xenobiotic drugs cause peroxisome proliferation in the liver. Defined functions for all these receptors, until recently, mainly concerned their ability to regulate energy balance, with PPARα being involved in β-oxidation pathways, and PPARγ in the differentiation of adipocytes. Little is known about the functions of PPARδ, though it is the most ubiquitously expressed. Since their discovery, PPARs have been shown to be expressed in monocytes/macrophages, the heart, vascular smooth muscle cells, endothelial cells, and in atherosclerotic lesions. Furthermore, PPARs can be activated by a vast number of compounds including synthetic drugs, of the clofibrate, and anti-diabetic thiazoldinedione classes, polyunsaturated fatty acids, and a number of eicosanoids, including prostaglandins, lipoxygenase products, and oxidized low density lipoprotein. This review will aim to introduce the field of PPAR nuclear hormone receptors, and discuss the discovery and actions of PPARs in the cardiovascular system, as well as the source of potential ligands. British Journal of Pharmacology (2000) 129, 823–834; doi:10.1038/sj.bjp.0703149
Article
Successful management of cardiovascular (CV) disease and associated metabolic syndromes, such as diabetes, is a major challenge to the clinician. Reducing CV risk factors, such as abnormal lipid profiles, insulin resistance or hypertension is the foundation of such therapy. A relatively new class of therapeutic agent, activators of peroxisome proliferator-activated receptors (PPAR), is poised to make a major impact with regard to several areas of risk factor management. However, there is growing evidence that PPAR agonists may also influence the CV system directly by modulating vessel wall function. These observations suggest that additional benefit, in the treatment of CV disease, may derive not only from the ability of agents to modify risk factors but also to influence directly the cellular mechanisms of disease within the vessel wall. A precedent for this dual action comes from examination of the effects of inhibitors of HMG CoA reductase (statins), where risk factor modulation is accompanied by direct actions on the vessel wall. In this review, we summarize the evidence suggesting that PPAR agonists may directly modulate vessel wall function, and that these may parallel those effects reported recently for the statins.
Article
Neuroprotective therapies for acute ischaemic stroke have yet to be realised despite the determined efforts of basic science and clinical investigators. Progressive elucidation of the complex pathophysiology involved in the ischaemic cascade has led to the development of numerous candidate interventions. Preliminary efficacy in animal models has repeatedly resulted in frustration after extensive clinical testing. Failure in the translation of results from animal models to humans implicates potential limitations of the current drug development process. Reflection on prior studies suggests possible flaws at several stages. Incorporation of standardised guidelines for preclinical testing of putative neuroprotective therapies and modification of clinical trial design, methodology and reporting may improve chances for success. The future of neuroprotection for stroke remains bright in spite of previous disappointments.
Article
Amyloid deposition within the brains of Alzheimer's Disease patients results in the activation of microglial cells and the induction of a local inflammatory response. The interaction of microglia or monocytes with beta-amyloid (A beta) fibrils elicits the activation a complex tyrosine kinase-based signal transduction cascade leading to stimulation of multiple independent signaling pathways and ultimately to changes in proinflammatory gene expression. The A beta-stimulated expression of proinflammatory genes in myeloid lineage cells is antagonized by the action of a family of ligand-activated nuclear hormone receptors, the peroxisome proliferator-activated receptors (PPARs). We report that THP-1 monocytes express predominantly PPAR gamma isoform and lower levels of PPAR alpha and PPAR delta isoforms. PPAR mRNA levels are not affected by differentiation of the cells into a macrophage phenotype, nor are they altered following exposure to the classical immune stimulus, lipopolysaccharide. Previous studies have found that PPAR gamma agonists act broadly to inhibit inflammatory responses. The present study explored the action of the PPAR alpha isoform and found that PPAR alpha agonists inhibited the A beta-stimulated expression of TNFalpha and IL-6 reporter genes in a dose-dependent manner. Moreover, the PPAR alpha agonist WY14643 inhibited macrophage differentiation and COX-2 gene expression. However, the PPAR alpha agonists failed to inhibit A beta-stimulated elaboration of neurotoxic factors by THP-1 cells. These findings demonstrate that PPAR alpha acts to suppress a diverse array of inflammatory responses in monocytes.