Progress in Neurobiology 90 (2010) 246–255
Contents lists available at ScienceDirect
Progress in Neurobiology journal homepage: www.elsevier.com/locate/pneurobio
Nitric oxide neurons and neurotransmission Steven R. Vincent * Department of Psychiatry & the Brain Research Centre, The University of British Columbia, Vancouver, BC, V6T 1Z3 Canada
A R T I C L E I N F O
A B S T R A C T
Article history: Received 30 November 2008 Received in revised form 22 April 2009 Accepted 9 October 2009
Nitric oxide was identified as a biological intercellular messenger just over 20 years ago, and its presence and potential importance in the nervous system was immediately noted. With the cloning of NO synthase and the physiological NO receptor soluble guanylyl cyclase, a variety of histochemical methods quickly led to a rather complete picture of where NO is produced and acts in the nervous system. However, the details regarding the subcellular localization of NO synthase and the identity of its molecular binding partners require further clarification. Although the hypothesis that calcium influx via activation of NMDA receptors is a key trigger for NO production has proven very popular and led to suggested roles for NO in synaptic plasticity, there is little direct evidence to support this notion. Instead, studies from the peripheral nervous system indicate a key role for voltage-sensitive calcium channels in regulating NO synthase activity. A similar mechanism may also be important in central neurons, and it remains an important task to identify the precise sources of calcium regulating NO production in specific NO neurons. Also, although cGMP production appears to mediate the physiological signaling by NO, the specific roles of cGMP-dependent ion channels, protein kinases and phosphodiesterases in mediating NO action remain to be determined. ß 2009 Elsevier Ltd. All rights reserved.
Keywords: Nitric oxide NADPH diaphorase Peripheral nervous system Central nervous system Sphenopalatine ganglion Enteric nervous system
Contents 1. 2.
3.
4.
Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . NO in the peripheral nervous system . . . . . . . . . . . . . . . . . . . . . . . . 2.1. NO in the innervation of the cranial blood vessels . . . . . . . . 2.2. The NO innervation of the penis . . . . . . . . . . . . . . . . . . . . . . 2.3. NO in the enteric nervous system and prevertebral ganglia. NO in the central nervous system. . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1. Nitric oxide in the cerebellum . . . . . . . . . . . . . . . . . . . . . . . . 3.2. Nitric oxide in the reticular activating system . . . . . . . . . . . 3.3. Nitric oxide in the forebrain. . . . . . . . . . . . . . . . . . . . . . . . . . Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1. Introduction Although nitric oxide has been known to have pharmacological actions for a great many years, the era of nitric oxide as an Abbreviations: AMPA, a-amino-3-hydroxyl-5-methyl-4-isoxazole-propionate; GABA, g-amino butyric acid; cAMP, cyclic adenosine monophosphate; cGMP, cyclic guanosine monophosphate; EDRF, endothelium-derived relaxing factor; LTS, low-threshold spike; NADPH, reduced nicotinamide adenine dinucleotide phosphate; NMDA, N-methyl-D-aspartate; NO, nitric oxide; ODQ, 1H-[1,2,4]oxadiazolo[4,3-a]quinoxalin-1-one; PSD-95, post-synaptic density protein of 95kD; VIP, vasoactive intestinal peptide. * Tel.: +1 604 822 7038; fax: +1 604 822 7981. E-mail address:
[email protected]. 0301-0082/$ – see front matter ß 2009 Elsevier Ltd. All rights reserved. doi:10.1016/j.pneurobio.2009.10.007
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
. . . . . . . . . . .
246 247 247 248 248 249 249 249 249 251 251
endogenous messenger molecule began just over 20 years ago. It is perhaps timely therefore to review the current research on the neurobiology of this still rather unusual molecule. In 1987, Palmer et al. (1987) reported that the biological activity of the endothelium-derived relaxing factor of Furchgott (Furchgott and Zawadzki, 1980) could be accounted for by nitric oxide release; a major discovery quickly confirmed by others (Ignarro et al., 1987). A synthetic pathway for nitric oxide from arginine was soon identified (Palmer et al., 1988; Schmidt et al., 1988) and the age of nitric oxide biology begun. That nitric oxide might also play a key role in the nervous system was indicated by the exciting discovery that NMDA treatment of cerebellar cultures resulted in the formation of an
S.R. Vincent / Progress in Neurobiology 90 (2010) 246–255
EDRF-like factor (Garthwaite et al., 1988). This was quickly shown to be due to the arginine-dependent synthesis of nitric oxide (Bredt and Snyder, 1989; Garthwaite et al., 1989). With the subsequent cloning and characterization of neuronal NO synthase it was shown to be a calmodulin-dependent enzyme (Bredt et al., 1991). Together, these observations led to the extremely attractive hypothesis, first enunciated by Garthwaite et al. (1988) that ‘‘influx of Ca2+ through NMDA-linked channels represents a likely trigger for EDRF production’’. Although this hypothesis has generated much research and speculation, it must be noted that in these studies of cerebellar slices and cultures, neuronal activity was never blocked, for example with tetrodotoxin, and thus the mechanisms mediating NMDA or glutamate induced NO formation in these preparations are not clear. There were hints in the past for this novel biosynthetic pathway. For example, Deguchi and Yoshioka already in 1982 identified L-arginine as an endogenous activator of guanylyl cyclase (Deguchi and Yoshioka, 1982). In addition, studies by Ratner et al. (1960) had demonstrated many years ago that the enzymes required for the conversion of citrulline to arginine were expressed in the brain. It now appears that the NO synthase expressing neurons often express high levels of arginine, citrulline and the enzymes arginosuccinate synthase and lyase (Arnt-Ramos et al., 1992; Braissant et al., 1999; Isayama et al., 1997; Nakamura, 1997; Nakamura et al., 1991; Nakata et al., 1991; Pasqualotto et al., 1991; Shuttleworth et al., 1995; Yu et al., 1997a,b). This arginine– citrulline cycle may allow NO synthase neurons to maintain adequate levels of arginine for NO synthesis. NO is an ancient messenger molecule. Various bacterial forms of NO synthase have been identified (Adak et al., 2002a,b; Pant et al., 2002). These appear similar to mammalian NO synthase, but lack an essential reductase domain to supply electrons for NO synthesis, and instead use other cellular redox partners (Gusarov et al., 2008). In metazoans, the NO synthase structure is highly conserved. A sea urchin NO synthase gene has been described (Cox et al., 2001), and genes for a neuronal NOS and a soluble guanylyl cyclase described in mollusks (Fujie et al., 2005; Matsuo et al., 2008). Insect NO synthase has been well characterized (Imamura et al., 2002; Ohtsuki et al., 2008; Regulski and Tully, 1995; Yuda et al., 1996) and a similar crustacean NO synthase described (Kim et al., 2004). One curiosity is the clear lack of NO synthase in the C. elegans genome (Morton et al., 1999). Why this particular creature should get along fine without this otherwise ubiquitous messenger is a puzzle; perhaps it is just to frustrate geneticists interested in this novel messenger. NO synthase has been described in all vertebrates examined, including fish (Cox et al., 2001), amphibians, reptiles, birds, and mammals including primates. Vertebrates express three distinct isoforms of NO synthase that are highly homologous, yet have distinct structures, regulation and distribution (Stuehr, 1999). The neuronal form was the first to be cloned and characterized (Bredt et al., 1991), and appears to be the isoform responsible for NO production in neurons. Indeed the overall distribution of NO synthase in the nervous system of vertebrate species is generally well conserved. Support for the idea of a role for NMDA-dependent calcium influx in the regulation of neuronal NO synthase came from the observation that neuronal NO synthase differs from other NO synthase isoforms in possessing a 230 residue N-terminal region containing a Class III PDZ domain (Tochio et al., 1999). NO synthase and PSD95 were found to interact in the yeast two-hybrid assay, and could be co-immunoprecipitated from cerebellar membranes (Brenman et al., 1996). This has led to the suggestion that PSD95 molecules may form a complex at synapses anchoring NO synthase in close proximity to NMDA receptors. Although such a NMDAPSD95-neuronal NO synthase complex has been generated in
247
transfected cell lines and shown to be functional (Ishii et al., 2006), it is not clear to what extent it exists or is functional in neurons. Indeed, a wide variety of other PDZ-dependent interactions with neuronal NO synthase have been reported (Chanrion et al., 2007; Firestein and Bredt, 1999; Hashida-Okumura et al., 1999; Jaffrey et al., 1998, 2002; Lemaire and McPherson, 2006; Manivet et al., 2000; Ort et al., 2000; Riefler and Firestein, 2001; Saitoh et al., 2004; Schuh et al., 2001; Stricker et al., 1997). An interesting recent project developed specific antibodies to the neuronal NO synthase b-finger domain mediating PDZ binding, and found that the overall distribution of immunolabeling obtained with the neuronal NO synthase b-finger antibody was very similar to that obtained with pan-neuronal NO synthase antibodies, indicating a lack of in situ PDZ-association and favoring a predominantly cytoplasmic distribution of NO synthase in various brain regions, including the hippocampus (Langnaese et al., 2007). Thus the extent to which neuronal NO synthase associates with various scaffolds and is targeted to specific subcellular domains remains an open question. The characterization of neuronal NO synthase led to the development of antibodies to this enzyme together with probes to detect the mRNA for this protein (Bredt et al., 1990). Furthermore, it was quickly shown that the NAPDH-diaphorase histochemical method provided a simple and specific technique with which to localize NO synthase (Hope et al., 1991). Together these various methods allowed the rapid determination of the distribution of NO synthase throughout the nervous system; the biochemical neuroanatomy of the NO system. These numerous studies have indicated that NO is produced in a wide variety of neurons of different types throughout the peripheral and central nervous systems (Aimi et al., 1991, 1993; Alm et al., 1993; Bredt et al., 1990; Burnett et al., 1993; Ceccatelli et al., 1994; Judas et al., 1999; Minami et al., 1994; Rodrigo et al., 1994; Vincent, 1994; Vincent and Kimura, 1992). These studies have also indicated that NO synthase can be found in the cell bodies, dendrites and axon terminals of neurons. 2. NO in the peripheral nervous system It might first be instructive to examine the peripheral nervous system in order to gain further insight into the regulation and actions of this novel messenger. In addition to being produced in endothelial cells as the EDRF by the endothelial form of NO synthase, NO is also a major messenger used by the peripheral nervous system (Vincent, 2000). Here it is produced by the neuronal form of the enzyme, although clearly, NMDA receptors do not mediate its formation. Instead, calcium influx through voltagegated ion channels, together with calcium release from intracellular stores are thought to trigger NO synthase activation. 2.1. NO in the innervation of the cranial blood vessels One area of great interest and importance is the innervation of the cerebral vasculature by NO neurons originating in the sphenopalatine ganglia (Fig. 1A–E) (Ceccatelli et al., 1994; Iadecola et al., 1993; Kimura et al., 1997; Minami et al., 1994; Morris et al., 1993; Nozaki et al., 1993; Suzuki et al., 1993; Uddman et al., 1999). In these neurons NO synthase coexists with the classical neurotransmitter acetylcholine and the peptide transmitter VIP (Fig. 1). Here, electrical stimulation of the nerves innervating the cerebral vessels produces an NO dependent relaxation that can be blocked by tetrodotoxin or an N-type calcium channel blocker (Toda et al., 1995). Indeed, N-type Ca2+ channels are the dominant voltage-dependent Ca2+ channels regulating Ca2+ influx during membrane depolarization of the NO synthase-positive sphenopalatine ganglion neurons (Liu et al., 2000) with R-type channels
248
S.R. Vincent / Progress in Neurobiology 90 (2010) 246–255
Fig. 1. Coexistence of choline acetyltransferase immunofluorescence (A), vasoactive intestinal polypeptide immunofluorescence (B) and NADPH diaphorase staining (C) of NO synthase in neurons in the rat sphenopalatine ganglion. NO synthase containing nerve fibers arising from these neurons heavily innervate the cerebral arteries (D and E) and the eye (F). Note the positive amacrine cells in the retina. The celiac ganglion receives a dense innervation (G) from the NADPH-diaphorase positive cells in the myenteric plexus. Following lesion of this input, a few of the positive fibers likely arising from sensory neurons and the sympathetic preganglionic neurons of the intermediolateral column remain (H).
contributing to a minor extent. Consistent with its role as the physiological NO receptor, inhibition of soluble guanylyl cyclase with ODQ (Garthwaite et al., 1995) prevented neurogenic relaxation (Gonzalez et al., 1997). 2.2. The NO innervation of the penis A similar situation appears to apply in the penis, where neurogenic relaxation was endothelium-independent and was inhibited by N-type Ca2+ channel antagonists (Okamura et al., 2001). Neurogenic NO activates soluble guanylyl cyclase and increases the generation of cyclic GMP in cavernosal muscle cells, resulting in the relaxation. NO synthase-positive postganglionic
parasympathetic nerves from the pelvic ganglia densely innervate cavernous tissues and penile arteries (Alm et al., 1993; Burnett et al., 1993; Dail et al., 1995; Keast, 1992; Schirar et al., 1994; Tamura et al., 1995). Many of these neurons also express VIP (Ehmke et al., 1995; Matsuda et al., 1996; Mizusawa et al., 2001; Tamura et al., 1995, 1997; Okamura et al., 2001). Again, inhibition of soluble guanylyl cyclase with ODQ prevented neurogenic relaxation (Recio et al., 1998; Simonsen et al., 2001). 2.3. NO in the enteric nervous system and prevertebral ganglia A large number of studies have demonstrated that NO synthase is expressed in populations of neurons throughout the gut (Alm
S.R. Vincent / Progress in Neurobiology 90 (2010) 246–255
et al., 1993; Anderson et al., 1995; Bredt et al., 1990; Ceccatelli et al., 1994; Costa et al., 1992; Dawson et al., 1991; Furness et al., 1994; Ward et al., 1992). It is found both in inhibitory motorneurons, and enteric interneurons (Yuan et al., 1995). This is consistent with work indicating that NO production may be responsible for relaxation of smooth muscle in the gut (Bayguinov and Sanders, 1993; Keef et al., 1993; Sanders and Ward, 1992; Shuttleworth et al., 1991). Muscle relaxation and NO release from enteric neurons can be evoked by field stimulation, and are blocked by tetrodotoxin or NO synthase inhibitors. NO synthesis in enteric neurons appears to be primarily dependent upon the influx of calcium via high-threshold calcium channels of the P, N and L variety (Kurjak et al., 2002). The prevertebral ganglia regulate a variety of visceral functions including digestive tract motility, secretion and absorption. The coeliac ganglion neurons receive a dense innervation of NO synthase fibers from the gut (Fig. 1G and H) that appears to regulate gut motility (Aimi et al., 1993; Furness and Anderson, 1994). This appears to be via the calcium-dependent production of NO and the formation of cGMP in the coeliac ganglion, where NO is released in response to splanchnic nerve stimulation (Quinson et al., 1999). Thus this again appears to be a useful model to examine NO production in response to activation of voltagesensitive calcium channels in nerve terminals. 3. NO in the central nervous system 3.1. Nitric oxide in the cerebellum The cerebellar cortex contains very high levels of NO synthase and NO production there appears to regulate functional hyperemia (Yang et al., 1999). NO synthase is expressed by local stellate, basket, and granule cells, but not Purkinje cells (Rodrigo et al., 1994; Vincent and Kimura, 1992). In an elegant study combining electrochemical NO detection with electrophysiology and imaging, Rancillac et al. (2006) confirmed that Purkinje cells do not express NO synthase. Instead they demonstrated that NMDA stimulation or direct depolarization of stellate cells in the molecular layer evoked NO release and local vasodilation. Interestingly, the NMDA evoked NO release was blocked by tetrodotoxin indicating that Ca2+ influx through the NMDA receptor may not be the trigger for NO production, and that spike firing of the stellate cells is required. High frequency stimulation of mossy fibers, which synapse on granule cells caused a significant NO release in the granular layer that was dependent on NMDA-receptor activation (Maffei et al., 2003). However, previous studies had shown that NO release in cerebellar slices in response to white matter stimulation presumably of parallel fibers was not affected by antagonists of NMDA receptors or other glutamate receptors (Kimura et al., 1998; Shibuki and Kimura, 1997) but was dramatically reduced by blockade of P/Q-type calcium channels (Shibuki and Kimura, 1997), and the cerebellar interneurons are known to express such channels (Stephens et al., 2001). In other words, a functional NMDA-receptor complex with NO synthase is not present in parallel fiber terminals (Shin and Linden, 2005). Instead, stellate and perhaps basket cells may mediate NMDA-dependent NO production. These interneurons express high levels of NO synthase (Fig. 2A) (Rodrigo et al., 1994; Vincent and Kimura, 1992) and NMDA receptors (Akazawa et al., 1994). Cerebellar NO also affects Purkinje cell function, in particular participating in the induction of the long-term depression of parallel fiber inputs (Casado et al., 2002; Ito, 1986; Lev-Ram et al., 1995; Levenes et al., 1998; Linden et al., 1995; Ogasawara et al., 2007; Shin and Linden, 2005). Purkinje cells express very high levels of soluble guanylyl cyclase (Bidmon et al., 2006; Ding et al., 2004; Gibb and Garthwaite, 2001; Giuili et al., 1994), type 1 cGMP-
249
dependent protein kinase (El-Husseini et al., 1999; Schlichter et al., 1980) and G-substrate (Endo et al., 1999; Hall et al., 1999; Nairn and Greengard, 1983), a phosphorylation-dependent inhibitor of protein phosphatase 1 and 2A. Thus, it appears that the NMDA receptor/NO cascade involved in cerebellar long-term depression is localized to interneurons rather than parallel fibers and that NO formed in these interneurons, perhaps via calcium influx through P/Q type calcium channels then acts in Purkinje cells dendrites to produce long-term depression (Shin and Linden, 2005). 3.2. Nitric oxide in the reticular activating system The cholinergic neurons of the laterodorsal and pedunculopontine tegmental nuclei express very high levels of NO synthase (Fig. 2C and D) and project to the thalamus (Aimi et al., 1991, 1993; Bredt et al., 1990; Dun et al., 1994; Minami et al., 1994; Rothe et al., 1999; Sugaya and McKinney, 1994; Usunoff et al., 1999; Vincent, 2000; Vincent and Kimura, 1992). At the ultrastructural level, this appears to be localized predominantly in the cytoplasm often in association with the endoplasmic reticulum (Rothe et al., 1999). Cholinergic neurons in the laterodorsal tegmental nucleus prominently express NMDA-receptor subunit immunoreactivity (Inglis and Semba, 1996) and are excited by glutamate through NMDA receptors (Sanchez and Leonard, 1996). In addition, R-type calcium channels may also play a key role in regulating NO production in these neurons (Kohlmeier and Leonard, 2006). Electrical stimulation or NMDA application evokes NO release from the somatodendritic region of these cells (Leonard et al., 2001). Local infusion of NMDA into the laterodorsal tegmental nucleus evokes NO and cGMP production, which in turn appears to modulate the release of noradrenaline from the adjacent locus ceruleus (Kodama and Koyama, 2006). In the thalamus, which is known to receive a massive innervation from the laterodorsal tegemental nucleus, NO is released following electrical stimulation of the laterodorsal tegmental nucleus and this NO formation in dependent upon action potential firing in these neurons (Miyazaki et al., 1996; Williams et al., 1997). Furthermore, endogenous NO production in the thalamus varies with behavioural state, being high during wake and REM sleep, and low during slow-wave sleep (Williams et al., 1997). Indeed, there is evidence that the level of NO production in these neurons may regulate the sleep–wake cycle (Hars, 1999). The mechanisms mediating NO action in the thalamus are not yet clear. Thalamic neurons express very high levels of type II cGMP-dependent protein kinase (El-Husseini et al., 1995, 1999; Vincent, 2000) and IRAG (Inositol 1,4,5-trisphosphate receptor associated cGMP kinase substrate (Geiselhoringer et al., 2004). NO in the thalamus appears to regulate cGMP-dependent protein kinase activity (El-Husseini et al., 1998). This signal cascade may regulate the firing pattern of thalamic neurons across behavioural states (Cudeiro et al., 2000; Pape and Mager, 1992; Shaw et al., 1999; Yang and Cox, 2008) or the processing of sensory information (Alexander et al., 2006; Cudeiro et al., 1994; Leamey et al., 2001). 3.3. Nitric oxide in the forebrain Another area in which NO neurons have received a great deal of attention is the forebrain. A population of NO synthase interneurons is present in the striatum, and similar cells are also present throughout the cortex. In the striatum, the NO synthase cells are aspiny interneurons that comprise only a few percent of the striatal neuronal population (Fig. 2B). These cells are GABAergic and usually also produce somatostatin and neuropeptide Y (Dawson et al., 1991; Figueredo-Cardenas et al., 1996; Kubota and
250
S.R. Vincent / Progress in Neurobiology 90 (2010) 246–255
Fig. 2. (A) Strongly stained NADPH-diaphorase positive stellate cells lie in the molecular layer (M) of the cerebellar cortex above the Purkinje cell layer (P) which is unstained, and the granule cell layer (G) which is weakly stained. In primary cultures of the rat striatum the intensely stained aspiny interneurons which express NADPH diaphorase are seen scattered among the unstained medium spiny neurons (B). The intensely stained NADPH diaphorase positive cells of the laterdorsal and adjacent pedunculopontine tegmental nuclei are shown (C). In situ hybridization using an oligonucleotide probe for neuronal NO synthase also strongly labels the cells of the laterodorsal tegmental nucleus (D).
Kawaguchi, 1994; Selden et al., 1994; Vincent et al., 1983). A similar population of cells is seen in the cortex (Dawson et al., 1991; Gonchar and Burkhalter, 1997; Higo et al., 2007; Judas et al., 1999; Kubota et al., 1994; Kubota and Kawaguchi, 1994; Smiley et al., 2000) and hippocampus (Fuentealba et al., 2008; Jinno et al., 1999; Jinno and Kosaka, 2004; Nomura et al., 1997; Seress et al., 2002; Valtschanoff et al., 1993). These forebrain interneurons all appear to have a common embryologic origin within the medial ganglionic eminence, from which they migrate tangentially to the striatum, hippocampus and cerebral cortex (Fogarty et al., 2007; Marin et al., 2000; Reid and Walsh, 2002). Recordings from identified NO synthase interneurons in the striatum indicated that these cells exhibited a unique phenotype characterized by calcium-dependent low-threshold spikes (LTS cells) (Kawaguchi, 1993; Kawaguchi et al., 1995). The majority of striatal NO synthase neurons had no detectable immunoreactivity for any of the AMPA receptor subunits (Bernard et al., 1997; Kwok et al., 1997), and there is not yet any compelling evidence for NMDA receptors on these aspiny interneurons. The other striatal cells, including the fast-spiking GABA interneurons and the cholinergic interneurons together with the medium spiny output
neurons receive direct glutamate input from the cortex (Fino et al., 2008), however, it is not yet clear if that is the case for the LTS NO synthase interneurons. In the cortex, this type of aspiny interneuron also generates low-threshold calcium spikes (Deuchars and Thomson, 1995; Goldberg et al., 2004; Kawaguchi, 1993; Kawaguchi et al., 1995). Goldberg et al. (2004) recorded identified cortical neurons from transgenic mice expressing green-fluorescent protein under the control of the somatostatin promoter. They made extremely interesting observations that these cells have AMPA receptors, but lack NMDA responses. Instead, low-threshold calcium spikes are evoked upon synaptic activation, such that the cells act as a ‘‘global single spiking unit’’ producing a global calcium signal throughout the entire dendritic tree of the cell. These cells exhibit ‘statedependent calcium signaling’ and fire in two ways, in burst mode, depolarization mediated by AMPA-mediated EPSPs activates Ttype calcium channels, which propagate through the neuron to trigger a global rise in calcium and a burst of action potentials. In tonic mode, when most T-channels are inactivated, AMPAmediated depolarizations evoke global action potentials (Goldberg et al., 2004). In contrast, in pyramidal cells and other types of
S.R. Vincent / Progress in Neurobiology 90 (2010) 246–255
cortical neurons, NMDA-receptor activation, combined with other conductances results in spatially confined calcium signals. One might speculate that the global calcium signals in the NOS interneurons might allow them to respond globally to discrete depolarizing synaptic inputs, releasing NO throughout the entire cortical area in which their processes ramify. The study of NO in the hippocampus has produced conflicting results. There has been such a plethora of publications suggesting a role for NO in various models of synaptic plasticity (Bohme et al., 1991, 1993; Bon et al., 1992; Bon and Garthwaite, 2001; Boulton et al., 1995; Doyle et al., 1996; Haley et al., 1992; Holscher, 1999; Hopper and Garthwaite, 2006; Mizutani et al., 1993; Musleh et al., 1993; Ramos et al., 2002; Schuman and Madison, 1991, 1994; Wu et al., 1997), that this concept has entered the textbooks as a dogma. However, there are numerous studies that do not support any such role for NO (Bannerman et al., 1994a,b; Chetkovich et al., 1993; Cummings et al., 1994; Tobin et al., 1995). Indeed, the presence of NO synthase in hippocampal pyramidal cells itself has been controversial, most reports initially indicating that hippocampal expression was confined to interneurons (Bredt et al., 1990; Rodrigo et al., 1994; Valtschanoff et al., 1993; Vincent and Kimura, 1992). However, after modifying fixation parameters and other variables, some sort of staining could be obtained by some groups within the CA1 pyramidal neurons (Blackshaw et al., 2003; Burette et al., 2002). Others have invoked the presence of the endothelial form of NO synthase in these cells (Dinerman et al., 1994; Kantor et al., 1996; O’Dell et al., 1994), but again this has not been seen by others, who instead report that endothelial NO synthase expression in the hippocampus is clearly confined to the endothelial cells (Blackshaw et al., 2003; Demas et al., 1999; Hashiguchi et al., 2004; Seidel et al., 1997; Stanarius et al., 1997; Topel et al., 1998). It is of interest that endogenous NO production in the hippocampus varies with behavioural state (Vincent et al., 1998), as noted previously in the thalamus (Williams et al., 1997). This NO may derive from the terminals of the septohippocampal neurons, many of which express NO synthase (Pasqualotto and Vincent, 1991), and contribute to the characteristic theta rhythm of neuronal activity in the hippocampus during active wake and REM sleep states (Simon et al., 2006; Vanderwolf et al., 1977). NO synthase containing interneurons in the hippocampus may also contribute, and these give rise to a dense network of fibers which ramify in close proximity to both pyramidal cell dendrites and arterioles, and may thus be able to regulate local blood flow in response to neuronal activity (Lovick et al., 1999). Forebrain regions express high levels of cGMP-stimulated phosphodiesterase (Repaske et al., 1993; Van Staveren et al., 2003). In particular, we have found that the medium spiny striatonigral neurons express this enzyme, and it appears to provide a mechanism whereby NO from aspiny interneurons, acting via cGMP can regulate dopamine-stimulated cAMP production (Lin et al., 2009). The role cGMP-regulated phosphodiesterases plays in mediating other NO actions deserves further study. 4. Conclusions Nitric oxide remains a fascinating but puzzling messenger in the nervous system. Although NADPH-diaphorase histochemistry combined with immunohistochemistry and in situ hybridization have clearly delineated the cells which synthesize NO and those which express its receptor, soluble guanylyl cyclase, much remains unknown regarding the significance of this signaling system in the central nervous system. The peripheral nervous system may suggest some clues. In particular the role of voltage-gated calcium channels in regulating the calcium influx that triggers NO synthesis
251
deserves further study. In addition, analogy with the regulation of cerebral and penile blood flow on the one hand, and gastric reflexes on the other, suggests that NO actions may be long-acting and state dependent. Indeed measurements of endogenous NO production in the brain support this sort of prolonged global action for NO in the nervous system. References Adak, S., Aulak, K.S., Stuehr, D.J., 2002a. Direct evidence for nitric oxide production by a nitric-oxide synthase-like protein from Bacillus subtilis. J. Biol. Chem. 277, 16167–16171. Adak, S., Bilwes, A.M., Panda, K., Hosfield, D., Aulak, K.S., McDonald, J.F., Tainer, J.A., Getzoff, E.D., Crane, B.R., Stuehr, D.J., 2002b. Cloning, expression, and characterization of a nitric oxide synthase protein from Deinococcus radiodurans. Proc. Natl. Acad. Sci. U.S.A. 99, 107–112. Aimi, Y., Fujimura, M., Vincent, S.R., Kimura, H., 1991. Localization of NADPHdiaphorase-containing neurons in sensory ganglia of the rat. J. Comp. Neurol. 306, 382–392. Aimi, Y., Kimura, H., Kinoshita, T., Minami, Y., Fujimura, M., Vincent, S.R., 1993. Histochemical localization of nitric oxide synthase in rat enteric nervous system. Neuroscience 53, 553–560. Akazawa, C., Shigemoto, R., Bessho, Y., Nakanishi, S., Mizuno, N., 1994. Differential expression of five N-methyl-D-aspartate receptor subunit mRNAs in the cerebellum of developing and adult rats. J. Comp. Neurol. 347, 150–160. Alexander, G.M., Kurukulasuriya, N.C., Mu, J., Godwin, D.W., 2006. Cortical feedback to the thalamus is selectively enhanced by nitric oxide. Neuroscience 142, 223– 234. Alm, P., Larsson, B., Ekblad, E., Sundler, F., Andersson, K.E., 1993. Immunohistochemical localization of peripheral nitric oxide synthase-containing nerves using antibodies raised against synthesized C- and N-terminal fragments of a cloned enzyme from rat brain. Acta Physiol. Scand. 148, 421–429. Anderson, C.R., Furness, J.B., Woodman, H.L., Edwards, S.L., Crack, P.J., Smith, A.I., 1995. Characterisation of neurons with nitric oxide synthase immunoreactivity that project to prevertebral ganglia. J. Auton. Nerv. Syst. 52, 107–116. Arnt-Ramos, L.R., O’Brien, W.E., Vincent, S.R., 1992. Immunohistochemical localization of argininosuccinate synthetase in the rat brain in relation to nitric oxide synthase-containing neurons. Neuroscience 51, 773–789. Bannerman, D.M., Chapman, P.F., Kelly, P.A., Butcher, S.P., Morris, R.G., 1994a. Inhibition of nitric oxide synthase does not impair spatial learning. J. Neurosci. 14, 7404–7414. Bannerman, D.M., Chapman, P.F., Kelly, P.A., Butcher, S.P., Morris, R.G., 1994b. Inhibition of nitric oxide synthase does not prevent the induction of long-term potentiation in vivo. J. Neurosci. 14, 7415–7425. Bayguinov, O., Sanders, K.M., 1993. Role of nitric oxide as an inhibitory neurotransmitter in the canine pyloric sphincter. Am. J. Physiol. 264, G975–G983. Bernard, V., Somogyi, P., Bolam, J.P., 1997. Cellular, subcellular, and subsynaptic distribution of AMPA-type glutamate receptor subunits in the neostriatum of the rat. J. Neurosci. 17, 819–833. Bidmon, H.J., Mohlberg, H., Habermann, G., Buse, E., Zilles, K., Behrends, S., 2006. Cerebellar localization of the NO-receptive soluble guanylyl cyclase subunitsalpha(2)/beta (1) in non-human primates. Cell Tissue Res. 326, 707–714. Blackshaw, S., Eliasson, M.J., Sawa, A., Watkins, C.C., Krug, D., Gupta, A., Arai, T., Ferrante, R.J., Snyder, S.H., 2003. Species, strain and developmental variations in hippocampal neuronal and endothelial nitric oxide synthase clarify discrepancies in nitric oxide-dependent synaptic plasticity. Neuroscience 119, 979–990. Bohme, G.A., Bon, C., Lemaire, M., Reibaud, M., Piot, O., Stutzmann, J.M., Doble, A., Blanchard, J.C., 1993. Altered synaptic plasticity and memory formation in nitric oxide synthase inhibitor-treated rats. Proc. Natl. Acad. Sci. U.S.A. 90, 9191– 9194. Bohme, G.A., Bon, C., Stutzmann, J.M., Doble, A., Blanchard, J.C., 1991. Possible involvement of nitric oxide in long-term potentiation. Eur. J. Pharmacol. 199, 379–381. Bon, C., Bohme, G.A., Doble, A., Stutzmann, J.M., Blanchard, J.C., 1992. A role for nitric oxide in long-term potentiation. Eur. J. Neurosci. 4, 420–424. Bon, C.L., Garthwaite, J., 2001. Exogenous nitric oxide causes potentiation of hippocampal synaptic transmission during low-frequency stimulation via the endogenous nitric oxide-cGMP pathway. Eur. J. Neurosci. 14, 585–594. Boulton, C.L., Southam, E., Garthwaite, J., 1995. Nitric oxide-dependent long-term potentiation is blocked by a specific inhibitor of soluble guanylyl cyclase. Neuroscience 69, 699–703. Braissant, O., Gotoh, T., Loup, M., Mori, M., Bachmann, C., 1999. L-arginine uptake, the citrulline-NO cycle and arginase II in the rat brain: an in situ hybridization study. Brain Res. Mol. Brain Res. 70, 231–241. Bredt, D.S., Hwang, P.M., Glatt, C.E., Lowenstein, C., Reed, R.R., Snyder, S.H., 1991. Cloned and expressed nitric oxide synthase structurally resembles cytochrome P-450 reductase. Nature 351, 714–718. Bredt, D.S., Hwang, P.M., Snyder, S.H., 1990. Localization of nitric oxide synthase indicating a neural role for nitric oxide. Nature 347, 768–770. Bredt, D.S., Snyder, S.H., 1989. Nitric oxide mediates glutamate-linked enhancement of cGMP levels in the cerebellum. Proc. Natl. Acad. Sci. U.S.A. 86, 9030– 9033. Brenman, J.E., Chao, D.S., Gee, S.H., McGee, A.W., Craven, S.E., Santillano, D.R., Wu, Z., Huang, F., Xia, H., Peters, M.F., Froehner, S.C., Bredt, D.S., 1996. Interaction of
252
S.R. Vincent / Progress in Neurobiology 90 (2010) 246–255
nitric oxide synthase with the postsynaptic density protein PSD-95 and alpha1syntrophin mediated by PDZ domains. Cell 84, 757–767. Burette, A., Zabel, U., Weinberg, R.J., Schmidt, H.H., Valtschanoff, J.G., 2002. Synaptic localization of nitric oxide synthase and soluble guanylyl cyclase in the hippocampus. J. Neurosci. 22, 8961–8970. Burnett, A.L., Tillman, S.L., Chang, T.S., Epstein, J.I., Lowenstein, C.J., Bredt, D.S., Snyder, S.H., Walsh, P.C., 1993. Immunohistochemical localization of nitric oxide synthase in the autonomic innervation of the human penis. J. Urol. 150, 73–76. Casado, M., Isope, P., Ascher, P., 2002. Involvement of presynaptic N-methyl-Daspartate receptors in cerebellar long-term depression. Neuron 33, 123–130. Ceccatelli, S., Lundberg, J.M., Zhang, X., Aman, K., Hokfelt, T., 1994. Immunohistochemical demonstration of nitric oxide synthase in the peripheral autonomic nervous system. Brain Res. 656, 381–395. Chanrion, B., Mannoury la Cour, C., Bertaso, F., Lerner-Natoli, M., Freissmuth, M., Millan, M.J., Bockaert, J., Marin, P., 2007. Physical interaction between the serotonin transporter and neuronal nitric oxide synthase underlies reciprocal modulation of their activity. Proc. Natl. Acad. Sci. U.S.A. 104, 8119–8124. Chetkovich, D.M., Klann, E., Sweatt, J.D., 1993. Nitric oxide synthase-independent long-term potentiation in area CA1 of hippocampus. Neuroreport 4, 919– 922. Costa, M., Furness, J.B., Pompolo, S., Brookes, S.J., Bornstein, J.C., Bredt, D.S., Snyder, S.H., 1992. Projections and chemical coding of neurons with immunoreactivity for nitric oxide synthase in the guinea-pig small intestine. Neurosci. Lett. 148, 121–125. Cox, R.L., Mariano, T., Heck, D.E., Laskin, J.D., Stegeman, J.J., 2001. Nitric oxide synthase sequences in the marine fish Stenotomus chrysops and the sea urchin Arbacia punctulata, and phylogenetic analysis of nitric oxide synthase calmodulin-binding domains. Comp. Biochem. Physiol. B Biochem. Mol. Biol. 130, 479–491. Cudeiro, J., Grieve, K.L., Rivadulla, C., Rodriguez, R., Martinez-Conde, S., Acuna, C., 1994. The role of nitric oxide in the transformation of visual information within the dorsal lateral geniculate nucleus of the cat. Neuropharmacology 33, 1413– 1418. Cudeiro, J., Rivadulla, C., Grieve, K.L., 2000. A possible role for nitric oxide at the sleep/wake interface. Sleep 23, 829–835. Cummings, J.A., Nicola, S.M., Malenka, R.C., 1994. Induction in the rat hippocampus of long-term potentiation (LTP) and long-term depression (LTD) in the presence of a nitric oxide synthase inhibitor. Neurosci. Lett. 176, 110–114. Dail, W.G., Barba, V., Leyba, L., Galindo, R., 1995. Neural and endothelial nitric oxide synthase activity in rat penile erectile tissue. Cell Tissue Res. 282, 109–116. Dawson, T.M., Bredt, D.S., Fotuhi, M., Hwang, P.M., Snyder, S.H., 1991. Nitric oxide synthase and neuronal NADPH diaphorase are identical in brain and peripheral tissues. Proc. Natl. Acad. Sci. U.S.A. 88, 7797–7801. Deguchi, T., Yoshioka, M., 1982. L-Arginine identified as an endogenous activator for soluble guanylate cyclase from neuroblastoma cells. J. Biol. Chem. 257, 10147– 10151. Demas, G.E., Kriegsfeld, L.J., Blackshaw, S., Huang, P., Gammie, S.C., Nelson, R.J., Snyder, S.H., 1999. Elimination of aggressive behavior in male mice lacking endothelial nitric oxide synthase. J. Neurosci. 19, RC30. Deuchars, J., Thomson, A.M., 1995. Innervation of burst firing spiny interneurons by pyramidal cells in deep layers of rat somatomotor cortex: paired intracellular recordings with biocytin filling. Neuroscience 69, 739–755. Dinerman, J.L., Dawson, T.M., Schell, M.J., Snowman, A., Snyder, S.H., 1994. Endothelial nitric oxide synthase localized to hippocampal pyramidal cells: implications for synaptic plasticity. Proc. Natl. Acad. Sci. U.S.A. 91, 4214–4218. Ding, J.D., Burette, A., Nedvetsky, P.I., Schmidt, H.H., Weinberg, R.J., 2004. Distribution of soluble guanylyl cyclase in the rat brain. J. Comp. Neurol. 472, 437–448. Doyle, C., Holscher, C., Rowan, M.J., Anwyl, R., 1996. The selective neuronal NO synthase inhibitor 7-nitro-indazole blocks both long-term potentiation and depotentiation of field EPSPs in rat hippocampal CA1 in vivo. J. Neurosci. 16, 418–424. Dun, N.J., Dun, S.L., Forstermann, U., 1994. Nitric oxide synthase immunoreactivity in rat pontine medullary neurons. Neuroscience 59, 429–445. Ehmke, H., Junemann, K.P., Mayer, B., Kummer, W., 1995. Nitric oxide synthase and vasoactive intestinal polypeptide colocalization in neurons innervating the human penile circulation. Int. J. Impot. Res. 7, 147–156. El-Husseini, A.E., Bladen, C., Vincent, S.R., 1995. Molecular characterization of a type II cyclic GMP-dependent protein kinase expressed in the rat brain. J. Neurochem. 64, 2814–2817. El-Husseini, A.E., Bladen, C., Williams, J.A., Reiner, P.B., Vincent, S.R., 1998. Nitric oxide regulates cyclic GMP-dependent protein kinase phosphorylation in rat brain. J. Neurochem. 71, 676–683. El-Husseini, A.E., Williams, J., Reiner, P.B., Pelech, S., Vincent, S.R., 1999. Localization of the cGMP-dependent protein kinases in relation to nitric oxide synthase in the brain. J. Chem. Neuroanat. 17, 45–55. Endo, S., Suzuki, M., Sumi, M., Nairn, A.C., Morita, R., Yamakawa, K., Greengard, P., Ito, M., 1999. Molecular identification of human G-substrate, a possible downstream component of the cGMP-dependent protein kinase cascade in cerebellar Purkinje cells. Proc. Natl. Acad. Sci. U.S.A. 96, 2467–2472. Figueredo-Cardenas, G., Morello, M., Sancesario, G., Bernardi, G., Reiner, A., 1996. Colocalization of somatostatin, neuropeptide Y, neuronal nitric oxide synthase and NADPH-diaphorase in striatal interneurons in rats. Brain Res. 735, 317–324. Fino, E., Deniau, J.M., Venance, L., 2008. Cell-specific spike-timing-dependent plasticity in GABAergic and cholinergic interneurons in corticostriatal rat brain slices. J. Physiol. 586, 265–282.
Firestein, B.L., Bredt, D.S., 1999. Interaction of neuronal nitric-oxide synthase and phosphofructokinase-M. J. Biol. Chem. 274, 10545–10550. Fogarty, M., Grist, M., Gelman, D., Marin, O., Pachnis, V., Kessaris, N., 2007. Spatial genetic patterning of the embryonic neuroepithelium generates GABAergic interneuron diversity in the adult cortex. J. Neurosci. 27, 10935–10946. Fuentealba, P., Begum, R., Capogna, M., Jinno, S., Marton, L.F., Csicsvari, J., Thomson, A., Somogyi, P., Klausberger, T., 2008. Ivy cells: a population of nitric-oxideproducing, slow-spiking GABAergic neurons and their involvement in hippocampal network activity. Neuron 57, 917–929. Fujie, S., Yamamoto, T., Murakami, J., Hatakeyama, D., Shiga, H., Suzuki, N., Ito, E., 2005. Nitric oxide synthase and soluble guanylyl cyclase underlying the modulation of electrical oscillations in a central olfactory organ. J. Neurobiol. 62, 14– 30. Furchgott, R.F., Zawadzki, J.V., 1980. The obligatory role of endothelial cells in the relaxation of arterial smooth muscle by acetylcholine. Nature 288, 373–376. Furness, J.B., Anderson, C.R., 1994. Origins of nerve terminals containing nitric oxide synthase in the guinea-pig coeliac ganglion. J. Auton. Nerv. Syst. 46, 47–54. Furness, J.B., Li, Z.S., Young, H.M., Forstermann, U., 1994. Nitric oxide synthase in the enteric nervous system of the guinea-pig: a quantitative description. Cell Tissue Res. 277, 139–149. Garthwaite, J., Charles, S.L., Chess-Williams, R., 1988. Endothelium-derived relaxing factor release on activation of NMDA receptors suggests role as intercellular messenger in the brain. Nature 336, 385–388. Garthwaite, J., Garthwaite, G., Palmer, R.M., Moncada, S., 1989. NMDA receptor activation induces nitric oxide synthesis from arginine in rat brain slices. Eur. J. Pharmacol. 172, 413–416. Garthwaite, J., Southam, E., Boulton, C.L., Nielsen, E.B., Schmidt, K., Mayer, B., 1995. Potent and selective inhibition of nitric oxide-sensitive guanylyl cyclase by 1H[1,2,4]oxadiazolo[4,3-a]quinoxalin-1-one. Mol. Pharmacol. 48, 184–188. Geiselhoringer, A., Gaisa, M., Hofmann, F., Schlossmann, J., 2004. Distribution of IRAG and cGKI-isoforms in murine tissues. FEBS Lett. 575, 19–22. Gibb, B.J., Garthwaite, J., 2001. Subunits of the nitric oxide receptor, soluble guanylyl cyclase, expressed in rat brain. Eur. J. Neurosci. 13, 539–544. Giuili, G., Luzi, A., Poyard, M., Guellaen, G., 1994. Expression of mouse brain soluble guanylyl cyclase and NO synthase during ontogeny. Brain Res. Dev. Brain Res. 81, 269–283. Goldberg, J.H., Lacefield, C.O., Yuste, R., 2004. Global dendritic calcium spikes in mouse layer 5 low threshold spiking interneurones: implications for control of pyramidal cell bursting. J. Physiol. 558, 465–478. Gonchar, Y., Burkhalter, A., 1997. Three distinct families of GABAergic neurons in rat visual cortex. Cereb. Cortex 7, 347–358. Gonzalez, C., Barroso, C., Martin, C., Gulbenkian, S., Estrada, C., 1997. Neuronal nitric oxide synthase activation by vasoactive intestinal peptide in bovine cerebral arteries. J. Cereb. Blood Flow Metab. 17, 977–984. Gusarov, I., Starodubtseva, M., Wang, Z.Q., McQuade, L., Lippard, S.J., Stuehr, D.J., Nudler, E., 2008. Bacterial nitric-oxide synthases operate without a dedicated redox partner. J. Biol. Chem. 283, 13140–13147. Haley, J.E., Wilcox, G.L., Chapman, P.F., 1992. The role of nitric oxide in hippocampal long-term potentiation. Neuron 8, 211–216. Hall, K.U., Collins, S.P., Gamm, D.M., Massa, E., DePaoli-Roach, A.A., Uhler, M.D., 1999. Phosphorylation-dependent inhibition of protein phosphatase-1 by Gsubstrate. A Purkinje cell substrate of the cyclic GMP-dependent protein kinase. J. Biol. Chem. 274, 3485–3495. Hars, B., 1999. Endogenous nitric oxide in the rat pons promotes sleep. Brain Res. 816, 209–219. Hashida-Okumura, A., Okumura, N., Iwamatsu, A., Buijs, R.M., Romijn, H.J., Nagai, K., 1999. Interaction of neuronal nitric-oxide synthase with alpha1-syntrophin in rat brain. J. Biol. Chem. 274, 11736–11741. Hashiguchi, A., Yano, S., Morioka, M., Hamada, J., Ushio, Y., Takeuchi, Y., Fukunaga, K., 2004. Up-regulation of endothelial nitric oxide synthase via phosphatidylinositol 3-kinase pathway contributes to ischemic tolerance in the CA1 subfield of gerbil hippocampus. J. Cereb. Blood Flow Metab. 24, 271–279. Higo, S., Udaka, N., Tamamaki, N., 2007. Long-range GABAergic projection neurons in the cat neocortex. J. Comp. Neurol. 503, 421–431. Holscher, C., 1999. Nitric oxide is required for expression of LTP that is induced by stimulation phase-locked with theta rhythm. Eur. J. Neurosci. 11, 335–343. Hope, B.T., Michael, G.J., Knigge, K.M., Vincent, S.R., 1991. Neuronal NADPH diaphorase is a nitric oxide synthase. Proc. Natl. Acad. Sci. U.S.A. 88, 2811–2814. Hopper, R.A., Garthwaite, J., 2006. Tonic and phasic nitric oxide signals in hippocampal long-term potentiation. J. Neurosci. 26, 11513–11521. Iadecola, C., Beitz, A.J., Renno, W., Xu, X., Mayer, B., Zhang, F., 1993. Nitric oxide synthase-containing neural processes on large cerebral arteries and cerebral microvessels. Brain Res. 606, 148–155. Ignarro, L.J., Buga, G.M., Wood, K.S., Byrns, R.E., Chaudhuri, G., 1987. Endotheliumderived relaxing factor produced and released from artery and vein is nitric oxide. Proc. Natl. Acad. Sci. U.S.A. 84, 9265–9269. Imamura, M., Yang, J., Yamakawa, M., 2002. cDNA cloning, characterization and gene expression of nitric oxide synthase from the silkworm, Bombyx mori. Insect Mol. Biol. 11, 257–265. Inglis, W.L., Semba, K., 1996. Colocalization of ionotropic glutamate receptor subunits with NADPH-diaphorase-containing neurons in the rat mesopontine tegmentum. J. Comp. Neurol. 368, 17–32. Isayama, H., Nakamura, H., Kanemaru, H., Kobayashi, K., Emson, P.C., Kawabuchi, M., Tashiro, N., 1997. Distribution and co-localization of nitric oxide synthase and argininosuccinate synthetase in the cat hypothalamus. Arch. Histol. Cytol. 60, 477–492.
S.R. Vincent / Progress in Neurobiology 90 (2010) 246–255 Ishii, H., Shibuya, K., Ohta, Y., Mukai, H., Uchino, S., Takata, N., Rose, J.A., Kawato, S., 2006. Enhancement of nitric oxide production by association of nitric oxide synthase with N-methyl-D-aspartate receptors via postsynaptic density 95 in genetically engineered Chinese hamster ovary cells: real-time fluorescence imaging using nitric oxide sensitive dye. J. Neurochem. 96, 1531–1539. Ito, M., 1986. Long-term depression as a memory process in the cerebellum. Neurosci. Res. 3, 531–539. Jaffrey, S.R., Benfenati, F., Snowman, A.M., Czernik, A.J., Snyder, S.H., 2002. Neuronal nitric-oxide synthase localization mediated by a ternary complex with synapsin and CAPON. Proc. Natl. Acad. Sci. U.S.A. 99, 3199–3204. Jaffrey, S.R., Snowman, A.M., Eliasson, M.J., Cohen, N.A., Snyder, S.H., 1998. CAPON: a protein associated with neuronal nitric oxide synthase that regulates its interactions with PSD95. Neuron 20, 115–124. Jinno, S., Aika, Y., Fukuda, T., Kosaka, T., 1999. Quantitative analysis of neuronal nitric oxide synthase-immunoreactive neurons in the mouse hippocampus with optical disector. J. Comp. Neurol. 410, 398–412. Jinno, S., Kosaka, T., 2004. Patterns of colocalization of neuronal nitric oxide synthase and somatostatin-like immunoreactivity in the mouse hippocampus: quantitative analysis with optical disector. Neuroscience 124, 797–808. Judas, M., Sestan, N., Kostovic, I., 1999. Nitrinergic neurons in the developing and adult human telencephalon: transient and permanent patterns of expression in comparison to other mammals. Microsc. Res. Tech. 45, 401–419. Kantor, D.B., Lanzrein, M., Stary, S.J., Sandoval, G.M., Smith, W.B., Sullivan, B.M., Davidson, N., Schuman, E.M., 1996. A role for endothelial NO synthase in LTP revealed by adenovirus-mediated inhibition and rescue. Science 274, 1744– 1748. Kawaguchi, Y., 1993. Physiological, morphological, and histochemical characterization of three classes of interneurons in rat neostriatum. J. Neurosci. 13, 4908– 4923. Kawaguchi, Y., Wilson, C.J., Augood, S.J., Emson, P.C., 1995. Striatal interneurones: chemical, physiological and morphological characterization. Trends Neurosci. 18, 527–535. Keast, J.R., 1992. A possible neural source of nitric oxide in the rat penis. Neurosci. Lett. 143, 69–73. Keef, K.D., Du, C., Ward, S.M., McGregor, B., Sanders, K.M., 1993. Enteric inhibitory neural regulation of human colonic circular muscle: role of nitric oxide. Gastroenterology 105, 1009–1016. Kim, H.W., Batista, L.A., Hoppes, J.L., Lee, K.J., Mykles, D.L., 2004. A crustacean nitric oxide synthase expressed in nerve ganglia, Y-organ, gill and gonad of the tropical land crab, Gecarcinus lateralis. J. Exp. Biol. 207, 2845–2857. Kimura, S., Uchiyama, S., Takahashi, H.E., Shibuki, K., 1998. cAMP-dependent longterm potentiation of nitric oxide release from cerebellar parallel fibers in rats. J. Neurosci. 18, 8551–8558. Kimura, T., Yu, J.G., Edvinsson, L., Lee, T.J., 1997. Cholinergic, nitric oxidergic innervation in cerebral arteries of the cat. Brain Res. 773, 117–124. Kodama, T., Koyama, Y., 2006. Nitric oxide from the laterodorsal tegmental neurons: its possible retrograde modulation on norepinephrine release from the axon terminal of the locus coeruleus neurons. Neuroscience 138, 245– 256. Kohlmeier, K.A., Leonard, C.S., 2006. Transmitter modulation of spike-evoked calcium transients in arousal related neurons: muscarinic inhibition of SNX482-sensitive calcium influx. Eur. J. Neurosci. 23, 1151–1162. Kubota, Y., Hattori, R., Yui, Y., 1994. Three distinct subpopulations of GABAergic neurons in rat frontal agranular cortex. Brain Res. 649, 159–173. Kubota, Y., Kawaguchi, Y., 1994. Three classes of GABAergic interneurons in neocortex and neostriatum. Jpn. J. Physiol. 44 (Suppl. 2), S145–S148. Kurjak, M., Sennefelder, A., Aigner, M., Schusdziarra, V., Allescher, H.D., 2002. Characterizing voltage-dependent Ca2+ channels coupled to VIP release and NO synthesis in enteric synaptosomes. Am. J. Physiol. 46, G1027–G1034. Kwok, K.H., Tse, Y.C., Wong, R.N., Yung, K.K., 1997. Cellular localization of GluR1, GluR2/3 and GluR4 glutamate receptor subunits in neurons of the rat neostriatum. Brain Res. 778, 43–55. Langnaese, K., Richter, K., Smalla, K.H., Krauss, M., Thomas, U., Wolf, G., Laube, G., 2007. Splice-isoform specific immunolocalization of neuronal nitric oxide synthase in mouse and rat brain reveals that the PDZ-complex-building nNOSalpha beta-finger is largely exposed to antibodies. Dev. Neurobiol. 67, 422– 437. Leamey, C.A., Ho-Pao, C.L., Sur, M., 2001. Disruption of retinogeniculate pattern formation by inhibition of soluble guanylyl cyclase. J. Neurosci. 21, 3871–3880. Lemaire, J.F., McPherson, P.S., 2006. Binding of Vac14 to neuronal nitric oxide synthase: characterisation of a new internal PDZ-recognition motif. FEBS Lett. 580, 6948–6954. Leonard, C.S., Michaelis, E.K., Mitchell, K.M., 2001. Activity-dependent nitric oxide concentration dynamics in the laterodorsal tegmental nucleus in vitro. J. Neurophysiol. 86, 2159–2172. Lev-Ram, V., Makings, L.R., Keitz, P.F., Kao, J.P., Tsien, R.Y., 1995. Long-term depression in cerebellar Purkinje neurons results from coincidence of nitric oxide and depolarization-induced Ca2+ transients. Neuron 15, 407–415. Levenes, C., Daniel, H., Crepel, F., 1998. Long-term depression of synaptic transmission in the cerebellum: cellular and molecular mechanisms revisited. Prog. Neurobiol. 55, 79–91. Lin, D.T.S., Fretier, P., Jiang, C., Vincent, S.R., 2009. Nitric oxide signaling via cGMPdependent phosphodiesterase in striatal neurons. Synapse, in press. Linden, D.J., Dawson, T.M., Dawson, V.L., 1995. An evaluation of the nitric oxide/ cGMP/cGMP-dependent protein kinase cascade in the induction of cerebellar long-term depression in culture. J. Neurosci. 15, 5098–5105.
253
Liu, J., Evans, M.S., Brewer, G.J., Lee, T.J., 2000. N-type Ca2+ channels in cultured rat sphenopalatine ganglion neurons: an immunohistochemical and electrophysiological study. J. Cereb. Blood Flow Metab. 20, 183–191. Lovick, T.A., Brown, L.A., Key, B.J., 1999. Neurovascular relationships in hippocampal slices: physiological and anatomical studies of mechanisms underlying flowmetabolism coupling in intraparenchymal microvessels. Neuroscience 92, 47– 60. Maffei, A., Prestori, F., Shibuki, K., Rossi, P., Taglietti, V., D’Angelo, E., 2003. NO enhances presynaptic currents during cerebellar mossy fiber-granule cell LTP. J. Neurophysiol. 90, 2478–2483. Manivet, P., Mouillet-Richard, S., Callebert, J., Nebigil, C.G., Maroteaux, L., Hosoda, S., Kellermann, O., Launay, J.M., 2000. PDZ-dependent activation of nitric-oxide synthases by the serotonin 2B receptor. J. Biol. Chem. 275, 9324–9331. Marin, O., Anderson, S.A., Rubenstein, J.L., 2000. Origin and molecular specification of striatal interneurons. J. Neurosci. 20, 6063–6076. Matsuda, H., Kusakabe, T., Kawakami, T., Takenaka, T., Sawada, H., Tsukuda, M., 1996. Coexistence of nitric oxide synthase and neuropeptides in the mouse vomeronasal organ demonstrated by a combination of double immunofluorescence labeling and a multiple dye filter. Brain Res. 712, 35–39. Matsuo, R., Misawa, K., Ito, E., 2008. Genomic structure of nitric oxide synthase in the terrestrial slug is highly conserved. Gene 415, 74–81. Minami, Y., Kimura, H., Aimi, Y., Vincent, S.R., 1994. Projections of nitric oxide synthase-containing fibers from the sphenopalatine ganglion to cerebral arteries in the rat. Neuroscience 60, 745–759. Miyazaki, M., Kayama, Y., Kihara, T., Kawasaki, K., Yamaguchi, E., Wada, Y., Ikeda, M., 1996. Possible release of nitric oxide from cholinergic axons in the thalamus by stimulation of the rat laterodorsal tegmental nucleus as measured with voltammetry. J. Chem. Neuroanat. 10, 203–207. Mizusawa, H., Hedlund, P., Hakansson, A., Alm, P., Andersson, K.E., 2001. Morphological and functional in vitro and in vivo characterization of the mouse corpus cavernosum. Br. J. Pharmacol. 132, 1333–1341. Mizutani, A., Saito, H., Abe, K., 1993. Involvement of nitric oxide in long-term potentiation in the dentate gyrus in vivo. Brain Res. 605, 309–311. Morris, R., Southam, E., Gittins, S.R., Garthwaite, J., 1993. NADPH-diaphorase staining in autonomic and somatic cranial ganglia of the rat. Neuroreport 4, 62–64. Morton, D.B., Hudson, M.L., Waters, E., O’Shea, M., 1999. Soluble guanylyl cyclases in Caenorhabditis elegans: NO is not the answer. Curr. Biol. 9, R546–R547. Musleh, W.Y., Shahi, K., Baudry, M., 1993. Further studies concerning the role of nitric oxide in LTP induction and maintenance. Synapse 13, 370–375. Nairn, A.C., Greengard, P., 1983. Cyclic GMP-dependent protein phosphorylation in mammalian brain. Fed. Proc. 42, 3107–3113. Nakamura, H., 1997. NADPH-diaphorase and cytosolic urea cycle enzymes in the rat spinal cord. J. Comp. Neurol. 385, 616–626. Nakamura, H., Saheki, T., Ichiki, H., Nakata, K., Nakagawa, S., 1991. Immunocytochemical localization of argininosuccinate synthetase in the rat brain. J. Comp. Neurol. 312, 652–679. Nakata, K., Mizuma, M., Nakagawa, S., Saheki, T., 1991. Distribution of argininosuccinate synthetase-like immunoreactive neurons in the rat myenteric plexus: a whole mount study. Okajimas Folia Anat. Jpn. 68, 209–212. Nomura, T., Fukuda, T., Aika, Y., Heizmann, C.W., Emson, P.C., Kobayashi, T., Kosaka, T., 1997. Distribution of nonprincipal neurons in the rat hippocampus, with special reference to their dorsoventral difference. Brain Res. 751, 64–80. Nozaki, K., Moskowitz, M.A., Maynard, K.I., Koketsu, N., Dawson, T.M., Bredt, D.S., Snyder, S.H., 1993. Possible origins and distribution of immunoreactive nitric oxide synthase-containing nerve fibers in cerebral arteries. J. Cereb. Blood Flow Metab. 13, 70–79. O’Dell, T.J., Huang, P.L., Dawson, T.M., Dinerman, J.L., Snyder, S.H., Kandel, E.R., Fishman, M.C., 1994. Endothelial NOS and the blockade of LTP by NOS inhibitors in mice lacking neuronal NOS. Science 265, 542–546. Ogasawara, H., Doi, T., Doya, K., Kawato, M., 2007. Nitric oxide regulates input specificity of long-term depression and context dependence of cerebellar learning. PLoS Comput. Biol. 3, e179. Ohtsuki, H., Yokoyama, J., Ohba, N., Ohmiya, Y., Kawata, M., 2008. Nitric oxide synthase (NOS) in the Japanese fireflies Luciola lateralis and Luciola cruciata. Arch. Insect. Biochem. Physiol. 69, 176–188. Okamura, T., Fujioka, H., Ayajiki, K., 2001. Effects of calcium antagonists on the nitrergic nerve function in canine corpus cavernosum. Jpn. J. Pharmacol. 87, 208–213. Ort, T., Maksimova, E., Dirkx, R., Kachinsky, A.M., Berghs, S., Froehner, S.C., Solimena, M., 2000. The receptor tyrosine phosphatase-like protein ICA512 binds the PDZ domains of beta2-syntrophin and nNOS in pancreatic beta-cells. Eur. J. Cell Biol. 79, 621–630. Palmer, R.M., Ashton, D.S., Moncada, S., 1988. Vascular endothelial cells synthesize nitric oxide from L-arginine. Nature 333, 664–666. Palmer, R.M., Ferrige, A.G., Moncada, S., 1987. Nitric oxide release accounts for the biological activity of endothelium-derived relaxing factor. Nature 327, 524– 526. Pant, K., Bilwes, A.M., Adak, S., Stuehr, D.J., Crane, B.R., 2002. Structure of a nitric oxide synthase heme protein from Bacillus subtilis. Biochemistry 41, 11071– 11079. Pape, H.C., Mager, R., 1992. Nitric oxide controls oscillatory activity in thalamocortical neurons. Neuron 9, 441–448. Pasqualotto, B.A., Hope, B.T., Vincent, S.R., 1991. Citrulline in the rat brain: immunohistochemistry and coexistence with NADPH-diaphorase. Neurosci. Lett. 128, 155–160.
254
S.R. Vincent / Progress in Neurobiology 90 (2010) 246–255
Pasqualotto, B.A., Vincent, S.R., 1991. Galanin and NADPH-diaphorase coexistence in cholinergic neurons of the rat basal forebrain. Brain Res. 551, 78–86. Quinson, N., Catalin, D., Niel, J.P., Miolan, J.P., 1999. Release of nitric oxide within the coeliac plexus is involved in the organization of a gastroduodenal inhibitory reflex in the rabbit. J. Physiol. 519 (Pt 1), 223–234. Ramos, A.J., Tagliaferro, P., Lopez-Costa, J.J., Lopez, E.M., Pecci Saavedra, J., Brusco, A., 2002. Neuronal and inducible nitric oxide synthase immunoreactivity following serotonin depletion. Brain Res. 958, 112–121. Rancillac, A., Rossier, J., Guille, M., Tong, X.K., Geoffroy, H., Amatore, C., Arbault, S., Hamel, E., Cauli, B., 2006. Glutamatergic control of microvascular tone by distinct GABA neurons in the cerebellum. J. Neurosci. 26, 6997–7006. Ratner, S., Morell, H., Carvalho, E., 1960. Enzymes of arginine metabolism in brain. Arch. Biochem. Biophys. 91, 280–289. Recio, P., Lopez, P.G., Hernandez, M., Prieto, D., Contreras, J., Garcia-Sacristan, A., 1998. Nitrergic relaxation of the horse corpus cavernosum. Role of Cgmp. Eur. J. Pharmacol. 351, 85–94. Regulski, M., Tully, T., 1995. Molecular and biochemical characterization of dNOS: a Drosophila Ca2+/calmodulin-dependent nitric oxide synthase. Proc. Natl. Acad. Sci. U.S.A. 92, 9072–9076. Reid, C.B., Walsh, C.A., 2002. Evidence of common progenitors and patterns of dispersion in rat striatum and cerebral cortex. J. Neurosci. 22, 4002–4014. Repaske, D.R., Corbin, J.G., Conti, M., Goy, M.F., 1993. A cyclic GMP-stimulated cyclic nucleotide phosphodiesterase gene is highly expressed in the limbic system of the rat brain. Neuroscience 56, 673–686. Riefler, G.M., Firestein, B.L., 2001. Binding of neuronal nitric-oxide synthase (nNOS) to carboxyl-terminal-binding protein (CtBP) changes the localization of CtBP from the nucleus to the cytosol: a novel function for targeting by the PDZ domain of nNOS. J. Biol. Chem. 276, 48262–48268. Rodrigo, J., Springall, D.R., Uttenthal, O., Bentura, M.L., Abadia-Molina, F., RiverosMoreno, V., Martinez-Murillo, R., Polak, J.M., Moncada, S., 1994. Localization of nitric oxide synthase in the adult rat brain. Philos. Trans. R. Soc. Lond. B Biol. Sci. 345, 175–221. Rothe, F., Huang, P.L., Wolf, G., 1999. Ultrastructural localization of neuronal nitric oxide synthase in the laterodorsal tegmental nucleus of wild-type and knockout mice. Neuroscience 94, 193–201. Saitoh, F., Tian, Q.B., Okano, A., Sakagami, H., Kondo, H., Suzuki, T., 2004. NIDD, a novel DHHC-containing protein, targets neuronal nitric-oxide synthase (nNOS) to the synaptic membrane through a PDZ-dependent interaction and regulates nNOS activity. J. Biol. Chem. 279, 29461–29468. Sanchez, R., Leonard, C.S., 1996. NMDA-receptor-mediated synaptic currents in guinea pig laterodorsal tegmental neurons in vitro. J. Neurophysiol. 76, 1101– 1111. Sanders, K.M., Ward, S.M., 1992. Nitric oxide as a mediator of nonadrenergic noncholinergic neurotransmission. Am. J. Physiol. 262, G379–G392. Schirar, A., Giuliano, F., Rampin, O., Rousseau, J.P., 1994. A large proportion of pelvic neurons innervating the corpora cavernosa of the rat penis exhibit NADPHdiaphorase activity. Cell Tissue Res. 278, 517–525. Schlichter, D.J., Detre, J.A., Aswad, D.W., Chehrazi, B., Greengard, P., 1980. Localization of cyclic GMP-dependent protein kinase and substrate in mammalian cerebellum. Proc. Natl. Acad. Sci. U.S.A. 77, 5537–5541. Schmidt, H.H., Klein, M.M., Niroomand, F., Bohme, E., 1988. Is arginine a physiological precursor of endothelium-derived nitric oxide? Eur. J. Pharmacol. 148, 293–295. Schuh, K., Uldrijan, S., Telkamp, M., Rothlein, N., Neyses, L., 2001. The plasmamembrane calmodulin-dependent calcium pump: a major regulator of nitric oxide synthase I. J. Cell. Biol. 155, 201–205. Schuman, E.M., Madison, D.V., 1991. A requirement for the intercellular messenger nitric oxide in long-term potentiation. Science 254, 1503–1506. Schuman, E.M., Madison, D.V., 1994. Locally distributed synaptic potentiation in the hippocampus. Science 263, 532–536. Seidel, B., Stanarius, A., Wolf, G., 1997. Differential expression of neuronal and endothelial nitric oxide synthase in blood vessels of the rat brain. Neurosci. Lett. 239, 109–112. Selden, N., Geula, C., Hersh, L., Mesulam, M.M., 1994. Human striatum: chemoarchitecture of the caudate nucleus, putamen and ventral striatum in health and Alzheimer’s disease. Neuroscience 60, 621–636. Seress, L., Abraham, H., Lin, H., Totterdell, S., 2002. Nitric oxide-containing pyramidal neurons of the subiculum innervate the CA1 area. Exp. Brain Res. 147, 38–44. Shaw, P.J., Charles, S.L., Salt, T.E., 1999. Actions of 8-bromo-cyclic-GMP on neurones in the rat thalamus in vivo and in vitro. Brain Res. 833, 272–277. Shibuki, K., Kimura, S., 1997. Dynamic properties of nitric oxide release from parallel fibres in rat cerebellar slices. J. Physiol. 498 (Pt 2), 443–452. Shin, J.H., Linden, D.J., 2005. An NMDA receptor/nitric oxide cascade is involved in cerebellar LTD but is not localized to the parallel fiber terminal. J. Neurophysiol. 94, 4281–4289. Shuttleworth, C.W., Burns, A.J., Ward, S.M., O’Brien, W.E., Sanders, K.M., 1995. Recycling of L-citrulline to sustain nitric oxide-dependent enteric neurotransmission. Neuroscience 68, 1295–1304. Shuttleworth, C.W., Murphy, R., Furness, J.B., 1991. Evidence that nitric oxide participates in non-adrendergic inhibitory transmission to intestinal muscle in the guinea-pig. Neurosci. Lett. 130, 77–80. Simon, A.P., Poindessous-Jazat, F., Dutar, P., Epelbaum, J., Bassant, M.H., 2006. Firing properties of anatomically identified neurons in the medial septum of anesthetized and unanesthetized restrained rats. J. Neurosci. 26, 9038–9046.
Simonsen, U., Contreras, J., Garcia-Sacristan, A., Martinez, A.C., 2001. Effect of sildenafil on non-adrenergic non-cholinergic neurotransmission in bovine penile small arteries. Eur. J. Pharmacol. 412, 155–169. Smiley, J.F., McGinnis, J.P., Javitt, D.C., 2000. Nitric oxide synthase interneurons in the monkey cerebral cortex are subsets of the somatostatin, neuropeptide Y, and calbindin cells. Brain Res. 863, 205–212. Stanarius, A., Topel, I., Schulz, S., Noack, H., Wolf, G., 1997. Immunocytochemistry of endothelial nitric oxide synthase in the rat brain: a light and electron microscopical study using the tyramide signal amplification technique. Acta Histochem. 99, 411–429. Stephens, G.J., Morris, N.P., Fyffe, R.E., Robertson, B., 2001. The Cav2.1/alpha1A (P/Qtype) voltage-dependent calcium channel mediates inhibitory neurotransmission onto mouse cerebellar Purkinje cells. Eur. J. Neurosci. 13, 1902–1912. Stricker, N.L., Christopherson, K.S., Yi, B.A., Schatz, P.J., Raab, R.W., Dawes, G., Bassett Jr., D.E., Bredt, D.S., Li, M., 1997. PDZ domain of neuronal nitric oxide synthase recognizes novel C-terminal peptide sequences. Nat. Biotechnol. 15, 336–342. Stuehr, D.J., 1999. Mammalian nitric oxide synthases. Biochim. Biophys. Acta 1411, 217–230. Sugaya, K., McKinney, M., 1994. Nitric oxide synthase gene expression in cholinergic neurons in the rat brain examined by combined immunocytochemistry and in situ hybridization histochemistry. Brain Res. Mol. Brain Res. 23, 111–125. Suzuki, N., Fukuuchi, Y., Koto, A., Naganuma, Y., Isozumi, K., Matsuoka, S., Gotoh, J., Shimizu, T., 1993. Cerebrovascular NADPH diaphorase-containing nerve fibers in the rat. Neurosci. Lett. 151, 1–3. Tamura, M., Kagawa, S., Kimura, K., Kawanishi, Y., Tsuruo, Y., Ishimura, K., 1995. Coexistence of nitric oxide synthase, tyrosine hydroxylase and vasoactive intestinal polypeptide in human penile tissue—a triple histochemical and immunohistochemical study. J. Urol. 153, 530–534. Tamura, M., Kagawa, S., Tsuruo, Y., Ishimura, K., Kimura, K., Kawanishi, Y., 1997. Localization of NADPH diaphorase and vasoactive intestinal polypeptide-containing neurons in the efferent pathway to the rat corpus cavernosum. Eur. Urol. 32, 100–104. Tobin, J.R., Gorman, L.K., Baxter, M.G., Traystman, R.J., 1995. Nitric oxide synthase inhibition does not impair visual or spatial discrimination learning. Brain Res. 694, 177–182. Tochio, H., Zhang, Q., Mandal, P., Li, M., Zhang, M., 1999. Solution structure of the extended neuronal nitric oxide synthase PDZ domain complexed with an associated peptide. Nat. Struct. Biol. 6, 417–421. Toda, N., Uchiyama, M., Okamura, T., 1995. Prejunctional modulation of nitroxidergic nerve function in canine cerebral arteries. Brain Res. 700, 213–218. Topel, I., Stanarius, A., Wolf, G., 1998. Distribution of the endothelial constitutive nitric oxide synthase in the developing rat brain: an immunohistochemical study. Brain Res. 788, 43–48. Uddman, R., Tajti, J., Moller, S., Sundler, F., Edvinsson, L., 1999. Neuronal messengers and peptide receptors in the human sphenopalatine and otic ganglia. Brain Res. 826, 193–199. Usunoff, K.G., Kharazia, V.N., Valtschanoff, J.G., Schmidt, H.H., Weinberg, R.J., 1999. Nitric oxide synthase-containing projections to the ventrobasal thalamus in the rat. Anat. Embryol. (Berl.) 200, 265–281. Valtschanoff, J.G., Weinberg, R.J., Kharazia, V.N., Nakane, M., Schmidt, H.H., 1993. Neurons in rat hippocampus that synthesize nitric oxide. J. Comp. Neurol. 331, 111–121. Van Staveren, W.C., Steinbusch, H.W., Markerink-Van Ittersum, M., Repaske, D.R., Goy, M.F., Kotera, J., Omori, K., Beavo, J.A., De Vente, J., 2003. mRNA expression patterns of the cGMP-hydrolyzing phosphodiesterases types 2, 5, and 9 during development of the rat brain. J. Comp. Neurol. 467, 566–580. Vanderwolf, C.H., Kramis, R., Robinson, T.E., 1977. Hippocampal electrical activity during waking behaviour and sleep: analyses using centrally acting drugs. Ciba Found Symp. 199–226. Vincent, S.R., 1994. Nitric oxide: a radical neurotransmitter in the central nervous system. Prog. Neurobiol. 42, 129–160. Vincent, S.R., 2000. The ascending reticular activating system—from aminergic neurons to nitric oxide. J. Chem. Neuroanat. 18, 23–30. Vincent, S.R., Johansson, O., Hokfelt, T., Skirboll, L., Elde, R.P., Terenius, L., Kimmel, J., Goldstein, M., 1983. NADPH-diaphorase: a selective histochemical marker for striatal neurons containing both somatostatin- and avian pancreatic polypeptide (APP)-like immunoreactivities. J. Comp. Neurol. 217, 252–263. Vincent, S.R., Kimura, H., 1992. Histochemical mapping of nitric oxide synthase in the rat brain. Neuroscience 46, 755–784. Vincent, S.R., Williams, J.A., Reiner, P.B., el-Husseini, A.E.-D., 1998. Monitoring neuronal NO release in vivo in cerebellum, thalamus and hippocampus. Prog. Brain Res. 118, 27–35. Ward, S.M., Xue, C., Shuttleworth, C.W., Bredt, D.S., Snyder, S.H., Sanders, K.M., 1992. NADPH diaphorase and nitric oxide synthase colocalization in enteric neurons of canine proximal colon. Am. J. Physiol. 263, G277–G284. Williams, J.A., Vincent, S.R., Reiner, P.B., 1997. Nitric oxide production in rat thalamus changes with behavioral state, local depolarization, and brainstem stimulation. J. Neurosci. 17, 420–427. Wu, J., Wang, Y., Rowan, M.J., Anwyl, R., 1997. Evidence for involvement of the neuronal isoform of nitric oxide synthase during induction of long-term potentiation and long-term depression in the rat dentate gyrus in vitro. Neuroscience 78, 393–398. Yang, G., Chen, G., Ebner, T.J., Iadecola, C., 1999. Nitric oxide is the predominant mediator of cerebellar hyperemia during somatosensory activation in rats. Am. J. Physiol. 277, R1760–R1770.
S.R. Vincent / Progress in Neurobiology 90 (2010) 246–255 Yang, S., Cox, C.L., 2008. Excitatory and anti-oscillatory actions of nitric oxide in thalamus. J. Physiol. 586, 3617–3628. Yu, J.G., Ishine, T., Kimura, T., O’Brien, W.E., Lee, T.J., 1997a. L-citrulline conversion to L-arginine in sphenopalatine ganglia and cerebral perivascular nerves in the pig. Am. J. Physiol. 273, H2192–H2199. Yu, J.G., O’Brien, W.E., Lee, T.J., 1997b. Morphologic evidence for L-citrulline conversion to L-arginine via the argininosuccinate pathway in porcine cerebral perivascular nerves. J. Cereb. Blood Flow Metab. 17, 884–893.
255
Yuan, S.Y., Bornstein, J.C., Furness, J.B., 1995. Pharmacological evidence that nitric oxide may be a retrograde messenger in the enteric nervous system. Br. J. Pharmacol. 114, 428–432. Yuda, M., Hirai, M., Miura, K., Matsumura, H., Ando, K., Chinzei, Y., 1996. cDNA cloning, expression and characterization of nitric-oxide synthase from the salivary glands of the blood-sucking insect Rhodnius prolixus. Eur. J. Biochem. 242, 807–812.