Neuropeptide Y as an Endogenous Antiepileptic, Neuroprotective and Pro-Neurogenic Peptide

Share Embed


Descripción

Recent Patents on CNS Drug Discovery, 2006, 1, 315-324

315

Neuropeptide Y as an Endogenous Antiepileptic, Neuroprotective and ProNeurogenic Peptide Sara Xapellia, Fabienne Agassea, Raquel Ferreiraa, Ana P. Silvaa,b and João O. Malvaa,c,* a

Center for Neuroscience and Cell Biology of Coimbra, 3004-517 Coimbra, Portugal, bInstitute of Pharmacology and Therapeutics, Faculty of Medicine, University of Coimbra, 3004-504 Coimbra, Portugal, cInstitute of Biochemistry, Faculty of Medicine, University of Coimbra, 3004-504 Coimbra, Portugal Received: September 22, 2006; Accepted: September 28, 2006; Revised: October 6, 2006

Abstract: Neuropeptide Y (NPY) is a small peptide important in cardiovascular physiology, feeding, anxiety, depression and epilepsy. In the hippocampus, NPY is mainly produced and released by GABAergic interneurons and inhibits glutamatergic neurotransmission in the excitatory tri-synaptic circuit. Under epileptic conditions, there is a robust overexpression of NPY and NPY receptors particularly in the granular and pyramidal cells, contributing to the tonic inhibition of glutamate release and consequently to control the spread of excitability into other brain structures. Recently, an important role was attributed to NPY in neuroprotection against excitotoxicity and in the modulation of neurogenesis. In the present review we discuss the potential relevance of NPY and NPY receptors in neuroprotection and neurogenesis, with implications for brain repair strategies. Recent patents describing new NPY receptor antagonists directed to treat obesity and cardiovascular disorders were published. However, the NPYergic system may also prove to be a good target for the treatment of pharmaco-resistant forms of temporal lobe epilepsy, by acting on hyperexcitability, neuronal death or brain repair. In order to achieve new NPY-based antiepileptic and brain repair strategies, selective NPY receptor agonists able to reach their targets in the epileptic brain must be developed in the near future.

Keywords: NPY, NPY receptors, glutamate, epilepsy, excitotoxicity, neuroprotection, neurogenesis, brain repair. 1. INTRODUCTION Epilepsy is a group of neurological disorders characterized by the occurrence of spontaneous non-evoked seizure activity, widely distributed in human populations, affecting people of all ages, gender and social groups with an incidence of about 1-3% [1]. Several pre- and post-natal events may contribute to the pathogenesis of epilepsy [2,3] and epileptic crises are usually revealed several years after the epileptogenic insult, following a latent period free of crises [4]. In spite of the devastating consequences of these diseases, current medical therapy is largely symptomatic and poorly controlled by antiepileptic drugs [1,5,6]. Temporal lobe epilepsy (TLE) is one of the most common epilepsy forms in the adult human [7]. TLE can be acquired following brain aggression like stroke, trauma, and neurodegenerative diseases [8] but in the majority of the clinical cases the origin of epilepsy is not certain. TLE is a drug-resistant type of human adult focal epilepsy characterized by the spontaneous occurrence of complex partial seizures [9] with a well restricted epileptic focus (partial) accompanied with impairment of consciousness (complex) [5,10]. Changes in neuronal network structure and physiology have been associated with epileptogenesis and subsequent seizure expression including neuronal loss, increased excita*Address correspondence to this author at the Institute of Biochemistry, Faculty of Medicine, University of Coimbra, 3004-504 Coimbra, Portugal; Tel: +351 239 833369; Fax: +351 239 822776; E-mail: [email protected] 1574-8898/06 $100.00+.00

tion, altered inhibition, circuitry reorganization and synapse abnormalities [8]. Patients with chronic TLE and animals subjected to chemo-convulsants display widespread neuronal cell death, axonal sprouting, enhanced neurogenesis and the activation and proliferation of astrocytes and resident microglia through the brain, gliosis and consequent hippocampal sclerosis [5,11-14]. TLE is characterized by several histological aberrations and functional recurrent excitatory circuits in the hippocampus [15]. In hippocampal tissues, from TLE patients [1622] and experimental models [16,23-28] an abnormal morphology of the axons of granule cells in the dentate gyrus (DG) i.e. sprouting of hippocampal mossy fibers can be observed. Mossy fiber sprouting is a pronounced expansion of the projections of the granule cell axons outside their normal targets into the supragranular layers of DG [16,17,19,25] and within the hilus [25,28]. The vast majority of newly formed asymmetric mossy fiber synapses terminates on dendritic spines of granule cells [29,30], suggesting the formation of new aberrant and recurrent excitatory synapses. Seizures have been documented in acute slices from kainic acid-treated epileptic rats with robust mossy fiber sprouting [31,32] and a positive correlation between mossy fiber sprouting and the frequency of spontaneous seizures has been reported in in vivo animal models [33,34]. in vitro Electrophysiological experiments conducted in hippocampal slices obtained from kainic acid-treated rats have shown that focal application of glutamate to the dentate molecular layer evokes excitatory postsynaptic currents (EPSCs) [35] or evokes excitatory postsynaptic potentials (EPSPs) [36] in granule cells distant from the application place. Taken together, these data support the idea that new © 2006 Bentham Science Publishers Ltd.

316

Recent Patents on CNS Drug Discovery, 2006, Vol. 1, No. 3

mossy fiber synapses are mainly excitatory and contribute to the overall excitability of the hippocampal circuits. 2. NPY AND NPY RECEPTORS IN TLE 2.1. NPY in TLE Since the discovery of substance P, an 11-aminoacid peptide found in the human hippocampus, neocortex and in the gastrointestinal tract, there has been a profound interest in discovering new neuropeptides and targeting their biological function. One of the most attractive molecules to emerge was neuropeptide Y (NPY). Peptide YY (PYY) and pancreatic polypeptide (PP), gut endocrine peptides integrate the rest of this peptide family [37]. NPY has a preponderant role in stress-related behaviors (such as anxiety and depression), feeding, cardiovascular and memory functions and in the control of seizure activity [38,39]. NPY receptor family has six different subtypes which in turn belong to the G-protein coupled receptor superfamily. Y1, Y2, and Y5 receptors have shown to be the most prominent in the hippocampal formation. While Y3 has so far been detected in the solitary tract nucleus, Y4 binds preferentially to PP and y6 is functional only in the mouse and rabbit [40]. NPY is the most abundant peptide in the Central Nervous System (CNS) preferentially expressed in GABAergic interneurons but it is also present in long projecting neurons [41-43]. Following high frequency neuronal firing (such as that observed during an epileptic seizure) NPY is an efficient regulator of excitatory neurotransmission [44-46]. Although NPY is distributed broadly throughout the nervous system, seizure-related changes in NPY and NPY receptors are seen mostly in brain areas such as the hippocampus involved in the initiation and propagation of epileptic discharges [47,48]. In human epileptic brain a pronounced overexpression of NPY was observed in interneurons [16,49,50]. Moreover, in animal models of epilepsy, kainic-acid injected or kindled rats, NPY mRNA and immunoreactivity (ir) increase in the hippocampus, amygdala, striatum and entorhinal cortex as well as in other extralimbic areas [51-60]. NPY overexpression follows a distinct time-course after seizure induction in different brain areas and in distinct subsets of neurons [48,51,53-55]. Strong NPY-ir persists in granule cells of epileptic brain [61] including the brain of pilocarpine-treated epileptic rats [62] for several months. Moreover, NPY gene expression is also increased after multiple electroconvulsive stimulations (ECS) supporting the hypothesis that NPY can play a central role as an endogenous antiepileptic agent [63]. Accordingly, significant increase in Y1, Y2, and Y 5 receptors mRNA were found in the DG after ECS. Moreover, a recent study suggests that the mechanisms coupling NPY receptor stimulation to G-protein activation can be augmented after repeated ECS [64]. In addition, anticonvulsant drugs given acutely after kainic acid injection prevent both seizures and NPY overexpression [53-55]. Furthermore, NPY administration (intracerebroventricularly; icv) inhibits hippocampal seizures in vivo in DG, subiculum and in CA3 pyramidal cells [65,66]. NPY-deficient mice occasionally develop mild spontaneous seizures, and exhibit markedly enhanced susceptibility to motor seizures induced by convulsant agents. Accordingly, NPY (icv) infusion before kainic acid administration prevents death in these animals [67,68]. Additionally, during kindling epileptogenesis, NPY knoc-

Malva et al.

kout mice (KO) had higher seizure severity scores and longer after discharge periods than the wild type mice [69]. Tu and collaborators [70] have shown that a Y2 receptor antagonist (BIIE0246) not only blocked the effects of added NPY but also enhanced recurrent mossy fiber synaptic transmission, the frequency of mEPSCs, and the magnitude of mossy fiber-evoked granule cell epileptiform activity, implying that spontaneous release of NPY from the recurrent mossy fiber pathway is sufficient to regulate synaptic transmission. Also, in an adult model of generalized genetic epilepsy (Genetic Absence Epilepsy Rats of Strasbourg-GAERS) NPY was capable of suppressing absence seizures [71]. Another study has shown that endogenous NPY prevents recurrence of experimental febrile seizures since this effect was abolished by an NPY antagonist. The entorhinal cortex is an important relay station between the hippocampus and other brain areas, providing input to the DG and CA3 from its layer II, and receives outputs from the CA1 and subiculum to layer III [72]. The entorhinal cortex contains local NPY-positive fibers deriving from both local and external sources [73]. in vitro Studies demonstrated anti-epileptiform effects of NPY in mouse CA3, CA1 and subiculum [74,75] and frontal cortex [76-78]. Other studies show that NPY reversibly shortens the after potential duration of spontaneous epileptiform discharges in the mouse entorhinal cortex of two different mouse strains [79]. 2.2. Involvement of NPY Receptors in TLE 2.2.1. Y1 Receptors The highest levels of Y1 receptor mRNA in the brain of several mammalian species are consistently seen in forebrain regions including the cerebral cortex, the hippocampal formation, and several amygdaloid, thalamic, and hypothalamic nuclei [80-82]. Within the rat hippocampus, high to moderate levels of Y1 receptor mRNA is restricted to the CA3, CA2 and CA1 pyramidal cell layers but data on Y1 receptor mRNA on DG are discrepant [80,82,83]. Seizures markedly increase the density of Y2 receptors in the DG whereas the density of Y1 receptors declines [84,85] suggesting that down-regulation of Y1 receptors is a mechanism for protection against recurrent epileptic events and hyperexcitability [59]. In human epileptic dentate molecular layer a decrease in Y1 receptor binding was observed [49] whereas in kindled rats, Y1 receptor binding and mRNA are significantly reduced in DG, CA1 and CA3 areas [86,87]. Moreover, Y 1 receptors seem to mediate weak excitatory actions of NPY in rats, since the Y1 receptor antagonist (BIBP3236) inhibits kainic acid-induced seizures and delays kindling epileptogenesis while the Y1 receptor agonist ([Leu32, Pro34]-NPY) and NPY itself blocks these effects suggesting a proconvulsive role for the Y1 receptor [88-91]. However, other evidences do not support a direct role of Y1 receptors to limit synaptic excitation since exogenous Y1 receptor-preferring agonist cannot inhibit mossy fiber-to-CA3 field excitatory postsynaptic potentials (fEPSP) [74] and in vitro seizure activity are largely insensitive to exogenous Y1 receptor-preferring agonist [75,92]. Furthermore, Y1 receptors have also been shown to inhibit glutamate release from rat hippocampal synapto-

NPY in Epileptic Brain Repair

Recent Patents on CNS Drug Discovery, 2006, Vol. 1, No. 3

somes [93] and to reduce seizure-like activity in neurons and hippocampal slices [76,94]. More recent studies support the proconvulsive role of Y 1 receptors and anticonvulsive role of Y2 receptors, since the recombinant adeno-associated viral vector overexpressing NPY (rAAV-NPY) led to a twofold reduction of seizures induced by kainic acid in wild type and Y1 receptor KO mice without affecting seizures in Y2 receptor KO [95].

the NPY-mediated inhibition of both glutamatergic transmission and of epileptiform discharges in two slice models of TLE [105]. Also, in vivo studies have demonstrated that intrahippocampal injections of Y2-preferring agonists inhibited seizures caused by intrahippocampal injection with kainate [105]. Moreover, a more recent study also showed that Y2 mediate the anti-excitatory actions of NPY, since the rAAV-mediated hippocampal overexpression was ineffective in Y2 receptor KO mice and the seizure-induced mortality rate increased in these KO animals [95]. These evidences suggest an endogenous neuroprotective effect of NPY mainly via Y2 receptor activation in epileptogenesis.

2.2.2. Y2 Receptors In the human brain, Y2 receptor mRNA is present in different cerebral cortical areas and also in the subcortical white matter [96,97]. In the hippocampal formation, the granular layer of the DG shows the highest signal of Y2 receptor mRNA [96,97] but high levels of the Y2 receptor mRNA are also found in the CA2 and CA3 regions [96,97].

2.2.3. Y5 Receptors In post-mortem human brain Y 5 receptor mRNA is found in several regions [106]. High levels of Y5 receptor mRNA are found in the amygdala, substantia nigra and hypothalamus [106]. In the hippocampus, moderate levels of Y5 mRNA are found in neurons of the CA2, CA3, and CA4 regions and DG with lower levels in CA1 [106]. A minor almost undetectable, Y 5 mRNA level is found in the cerebral cortex [106].

The co-localization of Y2 receptor mRNA with NPY mRNA in some neuronal cells supports pharmacological studies indicating a presynaptic localization of Y2 autoreceptors, able to inhibit the release of NPY and other neurotransmitters [93,97,98]. Accordingly, Y 2 receptors have been implicated in the anticonvulsant properties of NPY [40,59]. In human epileptic dentate hilus, Y 2 receptor binding increases [49] and NPY application suppresses epileptiform activity [99]. In kindled rats, Y2 receptor binding is upregulated in DG, down-regulated in CA3 and unchanged in CA1 area [86]. Moreover, Roder and collaborators [100] showed that Y2 receptor binding decreases in kainic acidtreated rats compared to control animals except in the hilus of DG where it remains elevated. High density of Y2 receptors is present at the Schaffer collateral terminals in the CA1 sector of the hippocampus where NPY exerts a potent inhibitory action on glutamate release [93,101,102]. In addition, up-regulated Y2 receptors on mossy fibers together with high expression of NPY in the same terminals may represent an additional crucial site for inhibition of glutamate release and consequently of epileptic activity [61,91]. Furthermore, Y 2 receptor KO mice show enhanced epileptic activity after intrahippocampal application of kainate [101,103] and activation of Y2 receptors suppresses seizure activity in hippocampal slices in vitro [76] and in vivo models [91]. In the CA3 area Y2 receptor activation has been shown to reduce both ictaform and interictaform activity and Y2 receptor-preferring agonist mimicked the inhibitory effect of NPY while Y2 receptor antagonist blocked the effect of both NPY and Y2 receptor agonist [78,92,104]. In fact, a recent study has shown that an Y 2 receptor-preferring agonist mimicked and the Y2 receptor-selective antagonist blocked,

The involvement of Y5 receptors in epilepsy has been a matter of debate. Kopp and collaborators [107] observed an up-regulation of NPY Y5 receptor mRNA after kindling epileptogenesis. In studies using levetiracetam as an anticonvulsant against kindled seizures an increase of the levels of Y 1- and particularly Y 5 receptors-ir was observed in DG. In CA1 and CA3 areas, levetiracetam increased Y1- and Y2 receptor-ir levels by approximately 20% while Y5 receptors-ir were increased by about 300% [86]. These results suggest that maintaining or elevating normal levels of Y5 receptors subtype may be of importance to the anticonvulsant and antiepileptogenic effect of levetiracetam [86]. On the contrary, a down-regulation of Y5 receptor in kainic acid-treated and kindled rats was observed by Bregola and collaborators [108]. Accordingly, in the hippocampus of kindled rats, Y5 receptor binding levels were lower than Y1 and Y2 binding levels [86]. NPY or Y2-preferring agonist had no effect on spontaneous epileptiform bursts elicited in zero-Mg bathing medium using acute hippocampal slices from Y5 receptor KO mice [75,109]. Interestingly, also Y5 receptor KO mice are more susceptible to seizures [75]. Accordingly, some in vivo studies have shown that NPY icv infusion is a powerful inhibitor of motor as well as electroencephalographic (EEG) seizures induced by kainic acid and that this effect was mediated via receptors with a pharmacological profile similar to the recently cloned rat Y5

Table 1 - NPY and NPY Receptor Changes in TLE TLE

Changes

317

NPY

Y1 receptor

Y2 receptor

Human

Up-regulation interneurons [16,49,50,60]

Down-regulation dentate molecular layer [49]

Up-regulation DG [49]

Animal models

Up-regulation hippocampus, amygdala, striatum and entorhinal cortex [51-59]

Down-regulation DG, CA1 and CA3 [86,87]

Up-regulation DG [60,86]

Y5 receptor

Up-regulation [107] Down-regulation [108]

318

Recent Patents on CNS Drug Discovery, 2006, Vol. 1, No. 3

receptor [110]. Another study revealed that NPY-mediated inhibition of fEPSC, evoked EPSC and evoked EPSC and spontaneous EPSC is absent in hippocampal slices from Y5 receptor KO mice [74]. Moreover, a Y5-preferring agonist was shown to completely suppress stimulus-train-induced bursting (STIB) in the CA3 region of hippocampal slices [78]. But, more recent studies have shown that the antiepileptic actions of NPY are mediated by Y2 receptors and not by Y 5 receptors [105]. In fact, El Bahh and collaborators [105] demonstrated that Y5 receptor-preferring agonists had small but significant effects in two slice models of TLE that were blocked by Y2 receptor antagonist but not by the Y5 receptor antagonists. Also, in vivo intrahippocampal injections of Y5 receptor-preferring agonists inhibited seizures, induced by kainic acid, but again the Y5 receptor antagonist did not block this effect. It is possible that Y5 receptor activation may regulate excitability in other brain regions, since recent evidences in rats, in vivo, indicate that the threshold for hippocampal rapid kindling can be reduced using an Y5 receptor agonist and that this effect is blocked by Y5 receptor antagonists [111]. 2.3. Modulation of Glutamate Release by NPY Glutamate is the principal excitatory neurotransmitter in the mammalian CNS. Within the brain, this excitatory neurotransmitter is synthesized by presynaptic neurons and stored in synaptic vesicles. Upon neuronal activity synaptic vesicles fuse with the plasma membrane and glutamate is released into the synaptic cleft [112]. As reviewed by Baraban and Tallent [113] neuropeptides play a pivotal role in neuronal excitability as they have a presynaptic site of action and in some cases they inhibit the release of glutamate. The modulation of postsynaptic excitatory neurotransmitter receptors mediated by peptides is rarely seen in experimental studies. Synaptic transmission at the CA3 and CA1 synapse was shown to be inhibited presynaptically by NPY in the rat hippocampus through the action of Y2 receptors that cause inhibition of voltage-dependent calcium channels [102]. The role of NPY and NPY receptors targeting K+ and Ca channels was described by Sun and colleagues [114] in rat thalamic neurons. In addition, our group has provided increasing evidences for a role of NPY in the inhibition of intracellular Ca2+ changes and glutamate release. Silva and collaborators [93] demonstrated in rat hippocampal synaptosomes, the presynaptic modulation of intracellular Ca2+ concentration ([Ca 2+]i) by activating Y1 and Y2 receptors. Y 2 receptor activation showed the highest inhibition of the total glutamate release in the whole hippocampus, sustaining a NPY inhibitory effect over evoked-glutamate release via Y2 receptors. Accordingly, in rod bipolar cell axon terminals D’Angelo and Brecha [115] also showed how NPY modulates presynaptic levels of Ca2+ supporting the involvement of Y2 in glutamate release. 2+

Y1 receptor is thought to act as an autoreceptor regulating the release of NPY. Wang [116] provided evidences that NPY acting on Y1 receptor could inhibit Ca2+ entry leading to the inhibition of evoked glutamate release in the nerve terminals of rat cerebral cortex. Y2 receptors are located presynaptically and are believed to have an inhibitory effect

Malva et al.

on calcium-mediated release of glutamate. Our group has also provided evidences for the functional interaction of Y1 and Y 2 and Y 2 and Y 5 receptors, suggesting the formation of oligomers, in the presynaptic inhibition of glutamate release. Moreover, the inhibitory effect of NPY was achieved mainly via Y 2 receptors [117]. Y5 receptors seem to share with Y2 a role in the modulation of hippocampal excitability since NPY is not able to inhibit glutamate-mediated synaptic transmission in the absence of Y 5 as seen in Y 5 receptor KO mice. Its role in the modulation of neuronal excitability remains controversial while it is possible that micromolar concentrations of NPY agonists used in vitro may not be selectively enough to assign receptor subtype specific roles [44,113]. The study of NPY system has gained particular relevance in pathological conditions characterized by neuronal hyperexcitability such as epilepsy. The hippocampus is associated to the generation and propagation of seizure activity. As a likely candidate to regulate excitatory transmission several studies on the antiepileptic actions of NPY and its receptors have been performed. Exogenously applied NPY has anticonvulsant properties and NPY KO mice are more susceptible to seizures [118]. Also, repeated electroconvulsive stimulations reduce significantly NPY specific binding sites in the rat hippocampus [119]. Kopp and collaborators [107] had already suggested the contribution of Y2 in the suppression of hyperexcitability. As reviewed by Vezzani and Sperk [59], NPY release and presynaptic Y2 receptors seem to increase in epileptic tissue. Stimulation of Y2 and Y5 receptors had previously been shown to reduce acute seizures while blocking Y1 receptor assured protection using the kainic acid injection model in the rat [120]. These evidences in the animal model, taken together with studies conducted in human epileptic hippocampal tissue, support a neuromodulator role for NPY in the restriction and propagation of seizures [99]. The NPY system acts as an endogenous neuroprotective mechanism as NPY immunoreactive fibers are increased in sclerotic hippocampus of TLE patients along with an upregulation of Y2 receptors and downregulation of Y1 receptors [49]. The extensive distribution of NPY in the CNS and colocalization with other neurotransmitters and modulators propose a neuromodulator role for this peptide. Knowledge about NPY and its receptor family may unquestionably represent an important tool in the treatment of any pathology where glutamate receptor dysfunction may occur. 2.4. NPY and Neuroprotection Against Excitotoxicity Seizures are the result of an imbalance between inhibitory and excitatory transmission in neuronal circuits. This imbalance toward hyperexcitability is accompanied by excessive glutamate release that may be deleterious to neuronal survival, causing excitotoxicity. So, it is conceivable that synaptic modulators able to decrease glutamate release can exert a neuroprotective action against glutamateinduced neuronal death. Excitotoxicity corresponds to the supraphysiological stimulation of glutamate receptors that after a prolonged activation leads to an increase in the intracellular Ca2+ concentration that is sufficiently high to trigger downstream processes resulting in cell death [121]. Therefore, the

NPY in Epileptic Brain Repair

receptors responsible for the excitotoxicity belong to the ionotropic glutamate receptors family. On the basis of pharmacological and electrophysiological properties this family has been divided into three subtypes: α-amino-3hydroxy-5methyl-isoxazole-4-propionate (AMPA), kainate and N-methyl-D-aspartate (NMDA) receptors. Thus, control of extracellular glutamate concentrations is a critical homeostatic function that can affect both neuronal signaling and neuronal survival [122]. Our group investigated the involvement of NPY receptors in mediating neuroprotection against excitotoxic insults (with AMPA or kainate) in organotypic cultures of rat hippocampal slices [117]. For dentate granule cells, selective activation of Y1, Y2, or Y5 receptors had a neuroprotective effect against AMPA-induced toxicity whereas only the activation of Y2 receptors was effective for CA1 pyramidal cells [117]. When the slice cultures were exposed to kainate, the CA3 pyramidal cells displayed significant degeneration and in this case the activation of Y1, Y2, and Y5 receptors was neuroprotective. For the kainic acid-induced degeneration of CA1 pyramidal cells, we again found that only the Y2 receptor activation is effective [117]. This is consistent with the findings that the selective activation of Y1 and Y5 receptors did not modulate glutamate release in CA1 [93]. NPY can also regulate excitotoxicity in other brain areas such as amygdala and cortex. NPY-ir increases locally in the rat somato-sensory cortex when the excitotoxic lesion with D,L-homocysteic acid is made in this area [123]. The amygdalar neurochemical changes including NPY-ir increase in the basolateral nucleus, were seen when the ipsilateral insular cortex was lesioned [123]. 3. NPY AND NPY RECEPTORS AS ANTIEPILEPTIC AND PATENT TARGETS ON DRUG DISCOVERY Studies concerning the alteration in the NPY system, in TLE, following treatment with anti-epileptic drugs are also important, since they can give more direct information about the mechanisms involved in the anti-epileptic effect of NPY. Anticonvulsant drugs thiopental or the excitatory amino acid antagonist (MK-801) given acutely after kainic acid injection prevent both seizures and the increase in NPY expression [55]. Also, Midgley and collaborators [124,125] reported that phencyclidine and MK-801 reduced the NPY levels. Similarly, daily injections of phenobarbital, starting 3 days after the initial status epilepticus, suppress the increase in NPY expression [126]. Kindling is associated with an upregulation of hippocampal NPY mRNA levels and pretreatment with the anticonvulsant levetiracetam markedly delays the progression of kindling [86,126] by inhibiting kindling-induced rise in NPY mRNA and increasing the expression of Y1- and particularly Y5-like receptors in all hippocampal subfields [86]. Other studies show that in seizure sensitive gerbil NPY-ir in the hippocampus was lower than NPY-ir in seizure resistant gerbil. Following treatment with the anticonvulsant vigabatrin, the number of NPY-ir neurons and NPY mRNA expression was increased in the hilus and in the hippocampus [127]. In contrast, zonisamide markedly elevated the density of NPY-ir fibers in the DG, an effect not observed in other hippocampal areas [127]. A recent study has also shown that chronic valproate administration to rats resulted in an increase in NPY mRNA

Recent Patents on CNS Drug Discovery, 2006, Vol. 1, No. 3

319

and protein expression in the nucleus reticularis thalami and hippocampus, but not in the neocortex [128]. Taken together, these data indicate that some anticonvulsants suppress convulsions in TLE and this effect is accompanied by increasing synthesis of NPY that may contribute to the inhibition of hippocampal hyperexcitability. As discussed above, an important hallmark of TLE is the circuitry reorganization of the DG [17-23]. Moreover, the implication of NPY in epilepsy also came from the observed increase in NPY levels in rat models of epilepsy as well as from hippocampi removed surgically from human patients with TLE [16,49-59]. This elevation could represent a compensatory mechanism in response to hyperexcitability, supporting thus NPY as an endogenous anticonvulsant agent. In addition, co-release of NPY may explain, in part, why stimulation of the mossy fibers in vitro [32,129-131] or the perforant path in vivo [132] usually does not evoke reverberating excitation in the dentate gyrus, even in the presence of robust recurrent mossy fibers growth. Thus, de novo expression of NPY in the mossy fibers pathway may protect the hippocampus from seizures that originate in or involve the entorhinal cortex [70]. Taken together these results suggest that NPY and NPY receptors may represent a promising antiepileptogenic drug target or may help in the development of new anticonvulsants. Recently, NPY and NPY receptors [133] have been important targets for drug discovery. Several patents were published describing new splice variants of the pancreatic polypeptide family, including NPY, [134], new NPY receptor agonists [135], and the development of new antagonists of NPY receptors [136-144]. These inventions were directed to treat several diseases, including: rhinitis [135], arthritis [140], hypertension and cardio/cerebrovascular diseases [134,135,139,140,142,143], renal dysfunction [139,140], respiratory diseases [139], dyslipidemia [134], diabetes [134,140,141,144], cancer [142], sexual and reproductive disorders [136], depression [136], schizophrenia [142], Alzheimer’s disease [142], epilepsy [136] and specially obesity and eating disorders [134,136-144] where Y5 receptors have been a preferential target for research [136-138,144]. 4. NPY AND NEUROGENESIS 4.1. NPY and Hippocampal Neurogenesis Neuronal death and synaptic reorganization are some of the features of temporal lobe epilepsy. So, at the long-term, new strategies to replace dead neurons and aberrant synaptic circuits are required for the proper cure of the epileptic hippocampus. Neurogenesis proceeds continuously in the DG of the hippocampus. The subgranular zone (SGZ) of the DG harbours precursor cells that proliferate to generate neurons. Neurons produced in the SGZ migrate locally to the granule cell layer where they achieve their differentiation into new granule neurons; in the rat about 10 000 new neurons are produced each day from DG progenitor cells [145]. Accordingly, pharmacological suppression of hippocampal neurogenesis decreases learning capacity in rodent, demonstrating the physiological relevance of the phenomena [146,147].

320

Recent Patents on CNS Drug Discovery, 2006, Vol. 1, No. 3

Hippocampal neurogenesis is under the tight control of factors that modulate SGZ neurogenesis. For instance, cystatin C, a fibroblast growth factor-2 (FGF-2) co-factor, and insulin-like growth factor-1 (IGF-1) are locally released and promote proliferation and neurogenesis [148,149]. In the rat DG, immature neurons secrete the stem cell-derived neural stem/progenitor cell supporting factor (SDNSF) that promotes progenitor cell survival and self renewal [150]. Sonic hedgehog, a factor vital for neural development, and the neurotransmitter acetylcholine promote proliferation in the hippocampus [151,152]. Astrocytes release factors, including pro-inflammatory cytonines, that stimulate hippocampal neurogenesis [153,154]. Moreover, the soluble factor neurogenesin-1 (Ng-1), an antagonist of bone morphogenetic protein ( BMP), is secreted by astrocytes and counteracts the neurogenic inhibitory effect of BMP-4 [155]. Recently, NPY has been shown to promote neurogenesis in the DG. DG interneurons are a source of NPY and DG cells express Y1 receptors [156]. In Y1 receptor KO mice, both proliferation of the progenitor cells and emergence of doublecortin positive neurons are reduced [157]. This study convincingly demonstrates that NPY supports neurogenesis in the DG. 4.2. NPY and Neurogenesis in the Olfactory Epithelium The olfactory epithelium (OE) is another, however less known, site of adult neurogenesis. Located in the nasal cavity, OE is a pseudostratified epithelium composed of sustentacular (or supporting) non-neuronal cells, neuronal sensory neurons called olfactory receptor neurons (ORNs) and basal cells. In the upper part of the epithelium - i.e. bordering the nasal fossa - olfactory receptor neurons present microvillosities bearing receptors for odour molecules. Odorant molecules carried into the nose, bind to receptors and trigger the activation of a signalling pathway involving G proteins, opening of cAMP ion gated channels and consequent depolarisation. Resulting action potentials are transmitted via axons of the olfactory receptor neurons that extend in the basal part of the epithelium and cross the cribriform plate to reach the olfactory bulb. ORNs are constantly regenerated. Indeed, basal cells have stem cell properties as they proliferate, generate olfactory receptor precursor neurons that develop into new ORNs [158]. Like in the DG of the hippocampus, ORNs production is modulated by local cues. Growth factors such as FGF-2, EGF, TGF-α, (EGF: epidermal growth factor; TGF: transforming growth factor) and IGF-I stimulate basal cells and/or olfactory receptor precursor neurons proliferation [159-160]. In turn, mature ORNs secrete the growth and differentiation factor 11 (GDF11), member of the TGF-β family, that limits the proliferation of olfactory receptor neuron precursors [161,162]. NPY also plays a role in OE neurogenesis. In the adult mice, sustentacular cells express NPY [163]. NPY KO mice present a significant reduction of the number of olfactory receptor precursor neurons as compared to wild type mice suggesting that NPY stimulate proliferation of the basal cells. The proliferative role of NPY has been shown to be mediated by Y1 receptors [164].

Malva et al.

4.3. NPY and Injury-Induced Neurogenesis Numerous factors are up-regulated in neurogenic niches upon injury and are able to promote injury-induced neurogenesis. For instance, the expression of FGF-2 is enhanced in the DG following seizure, stimulating DG neurogenesis [165]. Following olfactory bulb ablation, dying ORNs of the OE secrete leukaemia inhibitory factor (LIF) that stimulates the proliferation of olfactory receptor neuron precursors and by this way ORNs replacement [166]. However, some pieces of evidence suggest that NPY may play a crucial role into injury-induced neurogenesis. Indeed, following seizure activity, NPY is overexpressed in the hippocampus [60], and in parallel, neurogenesis is increased in the DG [167-169]. In depression disorders, neurogenesis is decreased in the DG. Antidepressants used in depression treatment, like lithium and electroconvulsive therapy, increase SGZ neurogenesis [170]. Decrease of NPY mRNA levels has been observed in rats from the strain Flinders sensitive line (FSL) which present depressive behaviour. Electroconvulsive treatments increase levels of NPY in the DG of FSL rats [171]. It is then tempting to speculate that there might be an association between increased neurogenesis and up-regulation of endogenous NPY upon antidepressive treatments. 5. CURRENT & FUTURE DEVELOPMENTS Important gaps concerning NPY and neurogenesis must be filled. Indeed, no data are available on the action of NPY on subventricular zone (SVZ), the major neurogenic niche in the rodent brain. Neuroblasts are produced in the SVZ during the whole lifespan, migrate towards the olfactory bulb where they differentiate into interneurons [172] improving odour discrimination and memory [173,174]. No study hitherto demonstrated modulation of SVZ neurogenesis by NPY, despite the wide expression of NPY in the rodent brain [175]. SVZ and DG neurogenic systems share structural similarities. Indeed, in both tissues, stem/progenitor cells have been identified as astrocytes [176,177]. In both areas, neurogenesis is controlled by similar cues. However highly speculative, NPY, might modulate SVZ neurogenesis, but to confirm this, a huge amount of work has to be done in this area. This work will be of interest as SVZ cells represent an ethical and reliable source of cells for brain repair purposes. NPY, as a modulator of neurogenesis, is of particular interest for the development of future cell-based therapies to replace dead neurons in TLE and major neurodegenerative diseases. ACKNOWLEDGMENTS This work was supported by the Foundation for Science and Technology, Portugal, and FEDER Grant SFRH/ BD/14163/2003, POCTI/SAU-FCF/58492/2004 and POCTI/ FCB/46804/2002. REFERENCES [1] [2]

Morimoto K, Fahnestock M, Racine RJ. Kindling and status epilepticus models of epilepsy: rewiring the brain. Prog Neurobiol 2004; 73(1): 1-60. Salazar AM, Jabbari B, Vance SC, et al. Epilepsy after penetrating head injury. I. Clinical correlates: a report of the Vietnam Head Injury Study. Neurology 1985; 35(10): 14061414.

NPY in Epileptic Brain Repair [3] [4] [5] [6] [7] [8] [9] [10]

[11] [12]

[13] [14] [15] [16] [17] [18] [19]

[20] [21]

[22]

[23] [24] [25]

[26]

Berg AT, Shinnar S. The risk of seizure recurrence following a first unprovoked seizure: a quantitative review. Neurology 1991; 41(7): 965-972. Bernardino L, Ferreira R, Cristóvão AJ, Sales F, Malva JO. Inflammation and neurogenesis in temporal lobe epilepsy. Curr Drug Targets CNS Neurol Disord 2005; 4(4): 349-360. McNamara JO. Cellular and molecular basis of epilepsy. J Neurosci 1994; 14(6): 3413-3425. Avoli M, D'Antuono M, Louvel J, et al . Network and pharmacological mechanisms leading to epileptiform synchronization in the limbic system in vitro. Prog Neurobiol 2002; 68(3): 167-207. Engel J. Jr. Classifications of the International League Against Epilepsy: time for reappraisal. Epilepsia 1998; 39(9): 1014-1017. Bausch SB. Axonal sprouting of GABAergic interneurons in temporal lobe epilepsy. Epilepsy Behav 2005; 7(3): 390-400. Engel J. Jr. Introduction to temporal lobe epilepsy. Epilepsy Res 1996; 26(1): 141-150. Proposal for revised clinical and electroencephalographic classification of epileptic seizures. From the Commission on Classification and Terminology of the International League Against Epilepsy. Epilepsia 1981; 22(4): 489-501. Mathern GW, Babb TL, Leite JP, et al. The pathogenic and progressive features of chronic human hippocampal epilepsy. Epilepsy Res 1996; 26(1): 151-161. Parent JM, Yu TW, Leibowitz RT, et al. Dentate granule cell neurogenesis is increased by seizures and contributes to aberrant network reorganization in the adult rat hippocampus. J Neurosci 1997; 17(10): 3727-3738. Jankowsky JL, Patterson PH. The role of cytokines and growth factors in seizures and their sequelae. Prog Neurobiol 2001; 63(2): 125-149. Ben-Ari Y. Limbic seizure and brain damage produced by kainic acid: mechanisms and relevance to human temporal lobe epilepsy. Neuroscience 1985; 14(2): 375-403. Koyama R, Ikegaya Y. Mossy fiber sprouting as a potential therapeutic target for epilepsy. Curr Neurovasc Res 2004; 1(1): 3-10. de Lanerolle NC, Kim JH, Robbins RJ, Spencer DD. Hippocampal interneuron loss and plasticity in human temporal lobe epilepsy. Brain Res 1989; 495(2): 387-395. Sutula T, Cascino G, Cavazos J, Parada I, Ramirez L. Mossy fiber synaptic reorganization in the epileptic human temporal lobe. Ann Neurol 1989; 26(3): 321-330. Houser CR. Granule cell dispersion in the dentate gyrus of humans with temporal lobe epilepsy. Brain Res 1990; 535(2): 195-204. Houser CR, Miyashiro JE, Swartz BE, et al. Altered patterns of dynorphin immunoreactivity suggest mossy fiber reorganization in human hippocampal epilepsy. J Neurosci 1990; 10(1): 267282. Babb TL, Kupfer WR, Pretorius JK, Crandall PH, Levesque MF. Synaptic reorganization by mossy fibers in human epileptic fascia dentata. Neuroscience 1991; 42(2): 351-363. Mathern GW, Pretorius JK, Babb TL, Quinn B. Unilateral hippocampal mossy fiber sprouting and bilateral asymmetric neuron loss with episodic postictal psychosis. J Neurosurg 1995; 82(2): 228-233. Mathern GW, Pretorius JK, Babb TL. Influence of the type of initial precipitating injury and at what age it occurs on course and outcome in patients with temporal lobe seizures. J Neurosurg 1995; 82(2): 220-227. Cronin J, Dudek FE. Chronic seizures and collateral sprouting of dentate mossy fibers after kainic acid treatment in rats. Brain Res 1988; 474(1): 181-184. Sutula T, He XX, Cavazos J, Scott G. Synaptic reorganization in the hippocampus induced by abnormal functional activity. Science 1988; 239(4844): 1147-1150. Sutula T, Zhang P, Lynch M, et al. Synaptic and axonal remodeling of mossy fibers in the hilus and supragranular region of the dentate gyrus in kainate-treated rats. J Comp Neurol 1998; 390(4): 578-594. Mello LE, Cavalheiro EA, Tan AM, et al. Circuit mechanisms of seizures in the pilocarpine model of chronic epilepsy: cell loss and mossy fiber sprouting. Epilepsia 1993; 34(6): 985-995.

Recent Patents on CNS Drug Discovery, 2006, Vol. 1, No. 3 [27] [28]

[29] [30] [31]

[32]

[33]

[34] [35] [36] [37] [38] [39] [40]

[41] [42] [43] [44] [45] [46] [47]

[48] [49] [50]

321

Kotti T, Riekkinen PJ Sr, Miettinen R. Characterization of target cells for aberrant mossy fiber collaterals in the dentate gyrus of epileptic rat. Exp Neurol 1997; 146(2): 323-330. Wenzel HJ, Woolley CS, Robbins CA, Schwartzkroin PA. Kainic acid-induced mossy fiber sprouting and synapse formation in the dentate gyrus of rats. Hippocampus 2000; 10(3): 244-260. Cavazos JE, Zhang P, Qazi R, Sutula TP. Ultrastructural features of sprouted mossy fiber synapses in kindled and kainic acidtreated rats. J Comp Neurol 2003; 458(3): 272-292. Buckmaster PS, Zhang GF, Yamawaki R. Axon sprouting in a model of temporal lobe epilepsy creates a predominantly excitatory feedback circuit. J Neurosci 2002; 22(15): 6650-6658. Wuarin JP, Dudek FE. Electrographic seizures and new recurrent excitatory circuits in the dentate gyrus of hippocampal slices from kainate-treated epileptic rats. J Neurosci 1996; 16(14): 4438-4448. Patrylo PR, Dudek FE. Physiological unmasking of new glutamatergic pathways in the dentate gyrus of hippocampal slices from kainate-induced epileptic rats. J Neurophysiol 1998; 79(1): 418-429. Gorter JA, van Vliet EA, Aronica E, Lopes da Silva FH. Progression of spontaneous seizures after status epilepticus is associated with mossy fibre sprouting and extensive bilateral loss of hilar parvalbumin and somatostatin-immunoreactive neurons. Eur J Neurosci 2001; 13(4): 657-669. Zhang X, Cui SS, Wallace AE, et al. Relations between brain pathology and temporal lobe epilepsy. J Neurosci 2002; 22(14): 6052-6061. Molnar P, Nadler JV. Mossy fiber-granule cell synapses in the normal and epileptic rat dentate gyrus studied with minimal laser photostimulation. J Neurophysiol 1999; 82(4): 1883-1894. Lynch M, Sutula T. Recurrent excitatory connectivity in the dentate gyrus of kindled and kainic acid-treated rats. J Neurophysiol 2000; 83(2): 693-704. Larhammar D. Evolution of neuropeptide Y, peptide YY and pancreatic polypeptide. Regulatory Peptides 1996; 62: 1-11. Thorsell A, Heilig M, Diverse functions of neuropeptide Y revealed using genetically modified animals. Neuropeptides 2002; 36(2-3): 182-193. Heilig M, The NPY system in stress, anxiety and depression. Neuropeptides 2004; 38: 213-224. Silva AP, Xapelli S., Grouzmann E, Cavadas C. The putative neuroprotective role of neuropeptide Y in the Central Nervous System. Curr Drug Targets CNS Neurol Disord 2005; 4(4): 331347. Chronwall BM, DiMaggio DA, Massari VJ, et al. The anatomy of neuropeptide-Y-containing neurons in rat brain. Neuroscience 1985; 15(4): 1159-1181. De Quidt ME, Emson PC. Neuropeptide Y in the adrenal gland: characterization, distribution and drug effects. Neuroscience 1986; 19(3): 1011-1022. De Quidt ME, Emson PC. Distribution of neuropeptide Y-like immunoreactivity in the rat central nervous system-II. Immunohistochemical analysis. Neuroscience 1986; 18(3): 545-618. Baraban SC. Neuropeptide Y and epilepsy: recent progress, prospects and controversies. Neuropeptides 2004; 38(4): 261265. Hokfelt T. Neuropeptides in perspective: the last ten years. Neuron 1991; 7(6): 867-879. Hokfelt T, Johansson O, Ljungdahl A, Lundberg JM, Schultzberg M. Peptidergic neurones. Nature 1980; 284(5756): 515-521. Schwarzer C, Williamson JM, Lothman EW, Vezzani A, Sperk G. Somatostatin, neuropeptide Y, neurokinin B and cholecystokinin immunoreactivity in two chronic models of temporal lobe epilepsy. Neuroscience 1995; 69(3): 831-845. Sperk G, Marksteiner J, Gruber B, et al . Functional changes in neuropeptide Y- and somatostatin-containing neurons induced by limbic seizures in the rat. Neuroscience 1992; 50(4): 831-846. Furtinger S, Pirker S, Czech T, et al. Plasticity of Y1 and Y2 receptors and neuropeptide Y fibers in patients with temporal lobe epilepsy. J Neurosci 2001; 21(15): 5804-5812. Mathern GW, Babb TL, Pretorius JK, Leite JP. Reactive synaptogenesis and neuron densities for neuropeptide Y, somatostatin, and glutamate decarboxylase immunoreactivity in

322

[51] [52]

[53]

[54] [55] [56] [57]

[58]

[59]

[60]

[61]

[62]

[63]

[64]

[65] [66] [67] [68] [69] [70]

Recent Patents on CNS Drug Discovery, 2006, Vol. 1, No. 3 the epileptogenic human fascia dentata. J Neurosci 1995; 15(5 Pt 2): 3990-4004. Bellmann R, Widmann R, Olenik C, et al. Enhanced rate of expression and biosynthesis of neuropeptide Y after kainic acidinduced seizures. J Neurochem 1991; 56(2): 525-530. Marksteiner J, Sperk G, Maas D. Differential increases in brain levels of neuropeptide Y and vasoactive intestinal polypeptide after kainic acid-induced seizures in the rat. Naunyn Schmiedebergs Arch Pharmacol 1989; 339(1-2): 173-177. Marksteiner J, Ortler M, Bellmann R, Sperk G. Neuropeptide Y biosynthesis is markedly induced in mossy fibers during temporal lobe epilepsy of the rat. Neurosci Lett 1990; 112(2-3): 143-148. Marksteiner J, Lassmann H, Saria A, et al. Neuropeptide Levels after Pentylenetetrazol Kindling in the Rat. Eur J Neurosci 1990; 2(1): 98-103. Marksteiner J, Prommegger R, Sperk G. Effect of anticonvulsant treatment on kainic acid-induced increases in peptide levels. Eur J Pharmacol 1990; 181(3): 241-246. Nagaki S, Fukamauchi F, Sakamoto Y, et al. Upregulation of brain somatostatin and neuropeptide Y following lidocaineinduced kindling in the rat. Brain Res 2000; 852(2): 470-474. Plata-Salaman CR, Ilyin SE, Turrin NP, et al. Kindling modulates the IL-1beta system, TNF-alpha, TGF-beta1, and neuropeptide mRNAs in specific brain regions. Brain Res Mol Brain Res 2000; 75(2): 248-258. Vezzani A, Schwarzer C, Lothman EW, Williamson J, Sperk G. Functional changes in somatostatin and neuropeptide Y containing neurons in the rat hippocampus in chronic models of limbic seizures. Epilepsy Res 1996; 26(1): 267-279. Vezzani A, Sperk G. Overexpression of NPY and Y2 receptors in epileptic brain tissue: an endogenous neuroprotective mechanism in temporal lobe epilepsy? Neuropeptides 2004; 38(4): 245-252. Silva AP, Xapelli S, Pinheiro PS, et al. Up-regulation of neuropeptide Y levels and modulation of glutamate release through neuropeptide Y receptors in the hippocampus of kainateinduced epileptic rats. J Neurochem 2005; 93(1): 163-170. Vezzani A, Ravizza T, Moneta D et al. Brain-derived neurotrophic factor immunoreactivity in the limbic system of rats after acute seizures and during spontaneous convulsions: temporal evolution of changes as compared to neuropeptide Y. Neuroscience 1999; 90(4): 1445-1461. Scharfman HE, Goodman JH, Sollas AL. Granule-like neurons at the hilar/CA3 border after status epilepticus and their synchrony with area CA3 pyramidal cells: functional implications of seizure-induced neurogenesis. J Neurosci 2000; 20(16): 61446158. Mikkelsen JD, Woldbye DP. Accumulated increase in neuropeptide Y and somatostatin gene expression of the rat in response to repeated electroconvulsive stimulation. J Psychiatr Res 2006; 40(2): 153-159. Christensen DZ, Olesen MV, Kristiansen H, Mikkelsen JD, Woldbye DP. Unaltered neuropeptide Y (NPY)-stimulated [(35)S]GTPgammaS binding suggests a net increase in NPY signalling after repeated electroconvulsive seizures in mice. J Neurosci Res 2006; Aug 28; [Epub ahead of print]. Klemp K, Woldbye DP. Repeated inhibitory effects of NPY on hippocampal CA3 seizures and wet dog shakes. Peptides 2001; 22(3): 523-527. Woldbye DP, Madsen TM, Larsen PJ, Mikkelsen JD, Bolwig TG. Neuropeptide Y inhibits hippocampal seizures and wet dog shakes. Brain Res 1996; 737(1-2): 162-168. Baraban SC, Hollopeter G, Erickson JC, Schwartzkroin PA, Palmiter RD. Knock-out mice reveal a critical antiepileptic role for neuropeptide Y. J Neurosci 1997; 17(23): 8927-8936. Erickson JC, Clegg KE, Palmiter RD. Sensitivity to leptin and susceptibility to seizures of mice lacking neuropeptide Y. Nature 1996; 381(6581): 415-421. Shannon HE, Yang L. Seizure susceptibility of neuropeptide-Y null mutant mice in amygdala kindling and chemical-induced seizure models. Epilepsy Res 2004; 61(1-3): 49-62. Tu B, Timofeeva O, Jiao Y, Nadler JV. Spontaneous release of neuropeptide Y tonically inhibits recurrent mossy fiber synaptic transmission in epileptic brain. J Neurosci 2005; 25(7): 17181729.

Malva et al. [71] [72] [73] [74]

[75] [76] [77] [78]

[79] [80]

[81] [82] [83]

[84] [85]

[86]

[87] [88] [89]

[90]

[91] [92]

Stroud LM, O'Brien TJ, Jupp B, Wallengren C, Morris MJ. Neuropeptide Y suppresses absence seizures in a genetic rat model. Brain Res 2005; 1033(2): 151-156. Witter MP, Wouterlood FG, Naber PA, Van Haeften T. Anatomical organization of the parahippocampal-hippocampal network. Ann N Y Acad Sci 2000; 911: 1-24. Kohler C. Cytochemical architecture of the entorhinal area. Adv Exp Med Biol 1986; 203: 83-98. Guo H, Castro PA, Palmiter RD, Baraban SC. Y5 receptors mediate neuropeptide Y actions at excitatory synapses in area CA3 of the mouse hippocampus. J Neurophysiol 2002; 87(1): 558-566. Marsh DJ, Baraban SC, Hollopeter G, Palmiter RD. Role of the Y5 neuropeptide Y receptor in limbic seizures. Proc Natl Acad Sci USA 1999; 96(23): 13518-13523. Bijak M. Neuropeptide Y suppresses epileptiform activity in rat frontal cortex and hippocampus in vitro via different NPY receptor subtypes. Neurosci Lett 1999; 268(3): 115-118. Bijak M. Neuropeptide Y reduces epileptiform discharges and excitatory synaptic transmission in rat frontal cortex in vitro. Neuroscience 2000; 96(3): 487-494. Ho MW, Beck-Sickinger AG, Colmers WF. Neuropeptide Y(5) receptors reduce synaptic excitation in proximal subiculum, but not epileptiform activity in rat hippocampal slices. J Neurophysiol 2000; 83(2): 723-734. Woldbye DP, Nanobashvili A, Husum H, Bolwig TG, Kokaia M. Neuropeptide Y inhibits in vitro epileptiform activity in the entorhinal cortex of mice. Neurosci Lett 2002; 333(2): 127-130. Kishi T, Aschkenasi CJ, Choi BJ, et al. Neuropeptide Y Y1 receptor mRNA in rodent brain: distribution and colocalization with melanocortin-4 receptor. J Comp Neurol 2005; 482(3): 217243. Mikkelsen JD, Larsen PJ. A high concentration of NPY (Y1)receptor mRNA-expressing cells in the rat arcuate nucleus. Neurosci Lett 1992; 148(1-2): 195-198. Parker RM, Herzog H. Regional distribution of Y-receptor subtype mRNAs in rat brain. Eur J Neurosci 1999; 11(4): 14311448. Larsen PJ, Sheikh SP, Jakobsen CR, Schwartz TW, Mikkelsen JD. Regional distribution of putative NPY Y1 receptors and neurons expressing Y1 mRNA in forebrain areas of the rat central nervous system. Eur J Neurosci 1993; 5(12): 1622-1637. Schwarzer C, Kofler N, Sperk G. Up-regulation of neuropeptide Y-Y2 receptors in an animal model of temporal lobe epilepsy. Mol Pharmacol 1998; 53(1): 6-13. Gobbi M, Gariboldi M, Piwko C, et al. Distinct changes in peptide YY binding to, and mRNA levels of, Y1 and Y2 receptors in the rat hippocampus associated with kindling epileptogenesis. J Neurochem 1998; 70(4): 1615-1622. Husum H, Bolwig TG, Sanchez C, Mathe AA, Hansen SL. Levetiracetam prevents changes in levels of brain-derived neurotrophic factor and neuropeptide Y mRNA and of Y1- and Y5-like receptors in the hippocampus of rats undergoing amygdala kindling: implications for antiepileptogenic and moodstabilizing properties. Epilepsy Behav 2004; 5(2): 204-215. Kofler N, Kirchmair E, Schwarzer C, Sperk G. Altered expression of NPY-Y1 receptors in kainic acid induced epilepsy in rats. Neurosci Lett 1997; 230(2): 129-132. Benmaamar R, Pham L, Marescaux C, Pedrazzini T, Depaulis A. Induced down-regulation of neuropeptide Y-Y1 receptors delays initiation of kindling. Eur J Neurosci 2003; 18(4): 768-774. Gariboldi M, Conti M, Cavaleri D, Samanin R, Vezzani A. Anticonvulsant properties of BIBP3226, a non-peptide selective antagonist at neuropeptide Y Y1 receptors. Eur J Neurosci 1998; 10(2): 757-759. Husum H, Gruber SH, Bolwig TG, Mathe AA. Extracellular levels of NPY in the dorsal hippocampus of freely moving rats are markedly elevated following a single electroconvulsive stimulation, irrespective of anticonvulsive Y1 receptor blockade. Neuropeptides 2002; 36(5): 363-369. Vezzani A, Sperk G, Colmers WF. Neuropeptide Y: emerging evidence for a functional role in seizure modulation. Trends Neurosci 1999; 22(1): 25-30. Klapstein GJ, Colmers WF. Neuropeptide Y suppresses epileptiform activity in rat hippocampus in vitro. J Neurophysiol 1997; 78(3): 1651-1661.

NPY in Epileptic Brain Repair [93]

[94] [95]

[96]

[97]

[98]

[99] [100]

[101]

[102]

[103]

[104]

[105]

[106]

[107]

[108] [109] [110] [111] [112]

Silva AP, Carvalho AP, Carvalho CM, Malva JO. Modulation of intracellular calcium changes and glutamate release by neuropeptide Y1 and Y2 receptors in the rat hippocampus: differential effects in CA1, CA3 and dentate gyrus. J Neurochem 2001; 79(2): 286-296. Reibel S, Nadi S, Benmaamar R, et al. Neuropeptide Y and epilepsy: varying effects according to seizure type and receptor activation. Peptides 2001; 22(3): 529-539. Lin EJ, Young D, Baer K, Herzog H, During MJ. Differential actions of NPY on seizure modulation via Y1 and Y2 receptors: evidence from receptor knockout mice. Epilepsia 2006; 47(4): 773-780. Caberlotto L, Fuxe K, Rimland JM, Sedvall G, Hurd YL. Regional distribution of neuropeptide Y Y2 receptor messenger RNA in the human post mortem brain. Neuroscience 1998; 86(1): 167-178. Caberlotto L, Fuxe K, Hurd YL. Characterization of NPY mRNA-expressing cells in the human brain: co-localization with Y2 but not Y1 mRNA in the cerebral cortex, hippocampus, amygdala, and striatum. J Chem Neuroanat 2000; 20(3-4): 327337. King PJ, Widdowson PS, Doods HN, Williams G. Regulation of neuropeptide Y release by neuropeptide Y receptor ligands and calcium channel antagonists in hypothalamic slices. J Neurochem 1999; 73(2): 641-646. Patrylo PR, van den Pol AN, Spencer DD, Williamson A. NPY inhibits glutamatergic excitation in the epileptic human dentate gyrus. J Neurophysiol 1999; 82(1): 478-483. Roder C, Schwarzer C, Vezzani A, et al. Autoradiographic analysis of neuropeptide Y receptor binding sites in the rat hippocampus after kainic acid-induced limbic seizures. Neuroscience 1996; 70(1): 47-55. Klapstein GJ, Colmers WF. 4-Aminopyridine and low Ca2+ differentiate presynaptic inhibition mediated by neuropeptide Y, baclofen and 2-chloroadenosine in rat hippocampal CA1 in vitro. Br J Pharmacol 1992; 105(2): 470-474. Qian J, Colmers WF, Saggau P. Inhibition of synaptic transmission by neuropeptide Y in rat hippocampal area CA1: modulation of presynaptic Ca2+ entry. J Neurosci 1997; 17(21): 8169-8177. McQuiston AR, Colmers WF. Neuropeptide Y2 receptors inhibit the frequency of spontaneous but not miniature EPSCs in CA3 pyramidal cells of rat hippocampus. J Neurophysiol 1996; 76(5): 3159-3168. El Bahh B, Cao JQ, Beck-Sickinger AG, Colmers WF. Blockade of neuropeptide Y(2) receptors and suppression of NPY's antiepileptic actions in the rat hippocampal slice by BIIE0246. Br J Pharmacol 2002; 136(4): 502-509. El Bahh B, Balosso S, Hamilton T, et al. The anti-epileptic actions of neuropeptide Y in the hippocampus are mediated by Y2 and not Y5 receptors. Eur J Neurosci 2005; 22(6): 14171430. Nichol KA, Morey A, Couzens MH, et al. Conservation of expression of neuropeptide Y5 receptor between human and rat hypothalamus and limbic regions suggests an integral role in central neuroendocrine control. J Neurosci 1999; 19(23): 1029510304. Kopp J, Nanobashvili A, Kokaia Z, Lindvall O, Hokfelt T. Differential regulation of mRNAs for neuropeptide Y and its receptor subtypes in widespread areas of the rat limbic system during kindling epileptogenesis. Brain Res Mol Brain Res 1999; 72(1): 17-29. Bregola G, Dumont Y, Fournier A, et al. Decreased levels of neuropeptide Y(5) receptor binding sites in two experimental models of epilepsy. Neuroscience 2000; 98(4): 697-703. Baraban SC. Antiepileptic actions of neuropeptide Y in the mouse hippocampus require Y5 receptors. Epilepsia 2002; 43 Suppl 5: 9-13. Woldbye DP, Larsen PJ, Mikkelsen JD, et al. Powerful inhibition of kainic acid seizures by neuropeptide Y via Y5-like receptors. Nat Med 1997; 3(7): 761-764. Benmaamar R, Richichi C, Gobbi M, et al. Neuropeptide Y Y5 receptors inhibit kindling acquisition in rats. Regul Pept 2005; 125(1-3): 79-83. Takamori S. VGLUTs: “Exciting” times for glutamatergic research? Neuroscience Research 2006; 55: 343-351.

Recent Patents on CNS Drug Discovery, 2006, Vol. 1, No. 3 [113] [114]

[115] [116]

[117]

[118] [119]

[120] [121] [122] [123] [124] [125] [126]

[127]

[128]

[129]

[130] [131] [132] [133] *[134] *[135] *[136]

323

Baraban S, Tallent MK. Interneuron diversity series: interneuronal neuropeptides - endogenous regulators of neuronal excitability. Trends Neurosci 2004; 27(3): 135-142. Sun Q-Q, Huguenard JR, Prince DA. Neuropeptide Y receptors differentially modulate G-protein-activated inwardly rectifying K+ channels and high-voltage-activated Ca2+ channels in rat thalamic neurons. J Physiol 2001; 531(1): 67-79. D’Angelo I, Brecha NC. Y 2 receptor expression and inhibition of voltage-dependent Ca2+ influx into rod bipolar cell terminals. Neuroscience 2004; 125: 1039-1049. Wang S-J. Activation of neuropeptide Y Y1 receptors inhibits glutamate release through reduction of voltage-dependent Ca2+ entry in the rat cerebral cortex nerve terminals: suppression of this inhibitory effect by the protein kinase C-dependent facilitatory pathway. Neuroscience 2005; 134: 987-1000. Silva AP, Pinheiro PS, Carvalho AP, et al. Activation of neuropeptide Y receptors is neuroprotective against excitotoxicity in organotypic hippocampal slice cultures. FASEB J 2003; 17(9): 1118-1120. Kaga T, Fujimiya M, Inui A. Emerging functions of neuropeptide Y Y2 receptors in the brain. Peptides 2001; 22: 501-506. Madsen TM, Greisen MH, Nielsen SM, Bolwig TG, Mikkelsen JD. Electroconvulsive stimuli enhance both neuropeptide Y receptor Y1 and Y2 messenger RNA expression and levels of binding in the rat hippocampus. Neuroscience 2000; 98(1): 3339. Vezzani A, Rizzi M, Conti M, Samanin S. Modulatory role of neuropeptides in seizures induced in rats by stimulation of glutamate receptors. J Nutr 2000; 130: 1046S-1048S. Choi DW, Koh JY, Peters S. Pharmacology of glutamate neurotoxicity in cortical cell culture: attenuation by NMDA antagonists. J Neurosci 1988; 8(1): 185-196. Swanson RA, Duan S. Regulation of glutamate transporter function. Neuroscientist 1999; 5: 280-282. Cheung RT, Cechetto DF. Neuropeptide changes following excitotoxic lesion of the insular cortex in rats. J Comp Neurol 1995; 362(4): 535-550. Midgley LP, Bush LG, Gibb JW, Hanson GR. Characterization of phencyclidine-induced effects on neuropeptide Y systems in the rat caudate-putamen. Brain Res 1992; 593(1): 89-96. Midgley LP, Bush LG, Gibb JW, Hanson GR. Differential regulation of neuropeptide Y systems in limbic structures of the rat. J Pharmacol Exp Ther 1993; 267(2): 707-713. Marksteiner J, Sperk G. Concomitant increase of somatostatin, neuropeptide Y and glutamate decarboxylase in the frontal cortex of rats with decreased seizure threshold. Neuroscience 1988; 26(2): 379-385. Kwak SE, Kim JE, Kim DS, et al. Differential effects of vigabatrin and zonisamide on the neuropeptide Y system in the hippocampus of seizure prone gerbil. Neuropeptides 2005; 39(5): 507-513. Brill J, Lee M, Zhao S, Fernald RD, Huguenard JR. Chronic valproic acid treatment triggers increased neuropeptide y expression and signaling in rat nucleus reticularis thalami. J Neurosci 2006; 26(25): 6813-6822. Cronin J, Obenaus A, Houser CR, Dudek FE. Electrophysiology of dentate granule cells after kainate-induced synaptic reorganization of the mossy fibers. Brain Res 1992; 573(2): 305310. Hardison JL, Okazaki MM, Nadler JV. Modest increase in extracellular potassium unmasks effect of recurrent mossy fiber growth. J Neurophysiol 2000; 84(5): 2380-2389. Tauck DL, Nadler JV. Evidence of functional mossy fiber sprouting in hippocampal formation of kainic acid-treated rats. J Neurosci 1985; 5(4): 1016-1022. Buckmaster PS, Dudek FE. Network properties of the dentate gyrus in epileptic rats with hilar neuron loss and granule cell axon reorganization. J Neurophysiol 1997; 77(5): 2685-2696. Redrobe JP, Dumont Y, St Pierre JA, Quirion R. Multiple receptors for neuropeptide Y in the hippocampus: putative roles in seizures and cognition. Brain Res 1999; 848(1-2): 153-166. Shemesh, R., Kliger, Y., Neville, L.F., Bernstein, J., Eshel, D.: US20060052301A1 (2006). Potter, E.: DE69731635T2 (2005). Marzabadi, M.R., Wong, W.C., Noble, S.A., Desai, M.N.: US20066989379 (2006).

324

Recent Patents on CNS Drug Discovery, 2006, Vol. 1, No. 3

[137] *[138] *[139] [140] [141] [142] [143] *[144] [145] [146] [147] [148] [149] [150]

[151] [152]

[153] [154]

[155]

[156]

[157] [158]

Stamford, A., Dong, Y., McCombie, S.W., Wu, Y.: US20060030588A1 (2006). Dax, L., McNally, J., Youngman, M.: DE60023148T2 (2006). Dow, R.L., Hammond, M.: EP1003744B1 (2005). Breu, V., Dautzenberg, F., Guerry, P., Nettekoven, M.H., Pflieger, P.: US20020052356A1 (2002). Greenlee, W.J., Huang, Y., Kelly, J., M., McCombie, S.W., Stamford, A.W., Wu, Y.: US20030114517A1 (2003). Otake, N., Ogino, Y., Kanatani, A.: WO05080348A1 (2005). Hammond, M.: DE69922918T2 (2006). Stamford, A.W., Boyle, C.D., Huang, Y.: EO1237875B1 (2005). Cameron HA, McKay, RD. Adult neurogenesis produces a large pool of new granule cells in the dentate gyrus. J Comp Neurol 2001; 435, 406-417. Evan Praag H, Schinder AF, Christie BR, Toni N, Palmer TD, Gage FH. Functional neurogenesis in the adult hippocampus. Nature 2002; 415: 1030-1034. Shors TJ, Miesegaes G, Beylin A, Zhao M, Rydel T, Gould E. Neurogenesis in the adult is involved in the formation of trace memories. Nature 2001; 410: 372-376. Taupin P, Ray J, Fischer WH, et al. FGF-2-responsive neural stem cell proliferation requires CCg, a novel autocrine/paracrine cofactor. Neuron 2000; 28: 385-397. Anderson MF, Aberg MA, Nilsson M, Eriksson PS. Insulin-like growth factor-I and neurogenesis in the adult mammalian brain. Brain Res. Dev. Brain Res. 2002; 134: 115-122. Toda H, Tsuji M, Nakano I, Kobuke K, Hayashi T, Kasahara H, Takahashi J, Mizoguchi A, Houtani T, Sugimoto T, Hashimoto N, Palmer TD, Honjo T, Tashiro K. Stem cell-derived neural stem/progenitor cell supporting factor is an autocrine/paracrine survival factor for adult neural stem/progenitor cells. J Biol Chem 2003; 278: 35491-35500. Lai K, Kaspar BK, Gage FH, Schaffer DV. Sonic hedgehog regulates adult neural progenitor proliferation in vitro and in vivo. Nat Neurosci 2003; 6: 21-27. Kotani S, Yamauchi T, Teramoto T, Ogura H. Pharmacological evidence of cholinergic involvement in adult hippocampal neurogenesis in rats. Neuroscience 2006; Aug 2; [Epub ahead of print]. Song H, Stevens CF, Gage FH. Astroglia induce neurogenesis from adult neural stem cells. Nature 2002; 417: 39-44. Barkho BZ, Song H, Aimone JB, Smrt RD, Kuwabara T, Nakashima K, Gage FH, Zhao X. Identification of astrocyteexpressed factors that modulate neural stem/progenitor cell differentiation. Stem Cells Dev. 2006; 15: 407-21. Ueki T, Tanaka M, Yamashita K, et al. A novel secretory factor, Neurogenesin-1, provides neurogenic environmental cues for neural stem cells in the adult hippocampus. J Neurosci 2003; 23: 11732-11740. Kopp J, Xu ZQ, Zhang X, et al. Expression of the neuropeptide Y Y1 receptor in the CNS of rat and of wild-type and Y1 receptor knock-out mice. Focus on immunohistochemical localization. Neuroscience 2002; 111: 443-532. Howell OW, Doyle K, Goodman JH, et al. Neuropeptide Y stimulates neuronal precursor proliferation in the post-natal and adult dentate gyrus. J Neurochem 2005; 93: 560-70. Calof AL, Mumm JS, Rim PC, Shou J. The neuronal stem cell of the olfactory epithelium. J Neurobiol 1998; 36: 190-205.

Malva et al. [159] [160] [161] [162] [163] [164] [165]

[166] [167] [168] [169] [170] [171]

[172]

[173]

[174]

[175] [176] [177]

Beites CL, Kawauchi S, Crocker CE, Calof AL. Identification and molecular regulation of neural stem cells in the olfactory epithelium. Exp Cell Res 2005; 306: 309-16. McCurdy RD, Feron F, McGrath JJ, Mackay-Sim A. Regulation of adult olfactory neurogenesis by insulin-like growth factor-I. Eur J Neurosci 2005; 22: 1581-8. Wu HH, Ivkovic S, Murray RC, et al. Autoregulation of neurogenesis by GDF11. Neuron 2003; 37: 197-207. Hastings NB, Gould E. Neurons inhibit neurogenesis. Nat Med 2003; 9: 264-66. Hansel DE, Eipper BA, Ronnett GV. Neuropeptide Y functions as a neuroproliferative factor. Nature 2001; 410: 940-44. Hansel DE, Eipper BA, Ronnett GV. Regulation of olfactory neurogenesis by amidated neuropeptides. J Neurosci Res 2001; 66: 1-7. Yoshimura S, Teramoto T, Whalen MJ, et al. FGF-2 regulates neurogenesis and degeneration in the dentate gyrus after traumatic brain injury in mice. J Clin Invest 2003; 112: 12021210. Bauer S, Rasika S, Mauduit C, et al. Leukemia inhibitory factor is a key signal for injury-induced neurogenesis in the adult mouse brain. J Neurosci 2003; 23: 1792-1803. Cha BH, Akman C, Silveira DC, Liu X, Holmes GL. Spontaneous recurrent seizure following status epilepticus enhances dentate gyrus neurogenesis. Brain Dev 2004; 26: 394-97. Ernst C, Christie BR. Temporally specific proliferation events are induced in the hippocampus following acute focal injury. J Neurosci Res 2006; 83: 349-61. Jessberger S, Romer B, Babu H, Kempermann G. Seizures induce proliferation and dispersion of doublecortin-positive hippocampal progenitor cells. Exp Neurol 2005; 196: 342-51. Dranovsky A, Hen R. Hippocampal neurogenesis: regulation by stress and antidepressants. Biol Psychiatry 2006; 59: 1136-43. Jimenez-Vasquez PA, Diaz-Cabiale Z, Caberlotto L, et al. Electroconvulsive stimuli selectively affect behavior and neuropeptide Y (NPY) and NPY Y(1) receptor gene expressions in hippocampus and hypothalamus of Flinders Sensitive Line rat model of depression. Eur Neuropsychopharmacol 2006; Aug 8; [Epub ahead of print]. Betarbet R, Zigova T, Bakay RA, Luskin MB. Dopaminergic and GABAergic interneurons of the olfactory bulb are derived from the neonatal subventricular zone. Int J Dev Neurosci 1996; 14: 921-930. Gheusi G, Cremer H, McLean H, Chazal G, Vincent JD, Lledo PM. Importance of newly generated neurons in the adult olfactory bulb for odor discrimination. Proc Natl Acad Sci USA 2000; 97: 1823-1828. Rochefort C, Gheusi G, Vincent JD, Lledo PM. Enriched odor exposure increases the number of newborn neurons in the adult olfactory bulb and improves odor memory. J Neurosci 2002; 22: 2679-2689. Danger JM, Tonon MC, Jenks BG, et al. localization in the central nervous system and neuroendocrine functions. Fundam Clin Pharmacol 1990; 4(3): 307-40. Doetsch F, Caillé I, Lim DA, Garcia-Verdugo JM, AlvarezBuylla A, Subventricular zone astrocytes are neural stem cells in the adult mammalian brain. Cell 1999; 97: 703-716. Seri B, Garcia-Verdugo JM, McEwen BS, Alvarez-Buylla A. Astrocytes give rise to new neurons in the adult mammalian hippocampus. J Neurosci 2001; 21: 7153-7160.

Lihat lebih banyak...

Comentarios

Copyright © 2017 DATOSPDF Inc.