30.11.2012 Views

Nitric Oxide Mediated Signal Transduction in Networks of Human ...

Nitric Oxide Mediated Signal Transduction in Networks of Human ...

Nitric Oxide Mediated Signal Transduction in Networks of Human ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Fehlerhafte Abgabe!<br />

Orig<strong>in</strong>al Artikel dürfen nicht mehr abgedruckt werden!<br />

Kumulative Dissertationen:<br />

Bei kumulativen Dissertationen sowie Habilitationen wird von der<br />

Bibliothek e<strong>in</strong>e Pdf-Datei <strong>in</strong>s Netz gestellt. Da <strong>in</strong> diesen Dateien die<br />

Orig<strong>in</strong>alarbeiten <strong>in</strong>tegriert s<strong>in</strong>d, wird die Veröffentlichung der<br />

kumulativen Dissertation und Habilitation im Netz der TiHo künftig<br />

nur mit Verweis auf die Literaturstellen der Orig<strong>in</strong>alarbeiten erfolgen.<br />

E<strong>in</strong>e Änderung der Promotions-, PhD- oder Habilitationsordnungen<br />

muss nicht erfolgen. …<br />

(Quelle: Auszug aus dem Protokoll* der hochschulöffentlichen<br />

Sitzung des Senates am 10.09.2009; 04. Sitzung; TOP 2 – Bericht des<br />

Präsidenten)<br />

01.06.2010 Hart<strong>in</strong>ger


<strong>Nitric</strong> <strong>Oxide</strong> <strong>Mediated</strong> <strong>Signal</strong> <strong>Transduction</strong> <strong>in</strong> <strong>Networks</strong> <strong>of</strong><br />

<strong>Human</strong> Model Neurons<br />

Thesis<br />

Submitted <strong>in</strong> partial fulfillment <strong>of</strong> the requirements for the degree<br />

Doctor <strong>of</strong> Philosophy (PhD)<br />

Division <strong>of</strong> Cell Biology, Institute <strong>of</strong> Physiology<br />

University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e Hannover<br />

and<br />

Center for Systems Neuroscience Hannover<br />

Awarded by University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e Hannover<br />

by<br />

Million Adane Tegenge<br />

born <strong>in</strong><br />

Nekemte / Ethiopia<br />

Hannover, Germany, 2010


Supervisor: Pr<strong>of</strong>. Dr. Gerd Bicker<br />

Division <strong>of</strong> Cell Biology, Institute <strong>of</strong> Physiology,<br />

University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e Hannover<br />

Advisory group: Pr<strong>of</strong>. Dr. Herbert Hildebrandt<br />

Department <strong>of</strong> Cellular Chemistry<br />

Hannover Medical School<br />

Pr<strong>of</strong>. Dr. Stephan Ste<strong>in</strong>lechner<br />

Department <strong>of</strong> Zoology<br />

University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e Hannover<br />

First evaluation: Pr<strong>of</strong>. Dr. Gerd Bicker<br />

Pr<strong>of</strong>. Dr. Herbert Hildebrandt<br />

Pr<strong>of</strong>. Dr. Stephan Ste<strong>in</strong>lechner<br />

`<br />

Second evaluation: PD Dr. Wolfgang Blenau<br />

Institute <strong>of</strong> Biochemisty and Biology<br />

University <strong>of</strong> Potsdam, Germany<br />

Date <strong>of</strong> f<strong>in</strong>al exam<strong>in</strong>ation: April 09, 2010<br />

The present work was supported by grants from the DFG (FG 1103; BI 262/16-1) and BMBF<br />

(0313732; 013925D) to G. Bicker. M.A. Tegenge received a Georg-Christoph-Lichtenberg<br />

scholarship from the M<strong>in</strong>istry for Science and Culture <strong>of</strong> Lower Saxony.


Dedicated to Lydia and Betania


Publications:<br />

Tegenge, M. A., and Bicker, G. (2009). <strong>Nitric</strong> oxide and cGMP signal transduction positively<br />

regulates the motility <strong>of</strong> human neuronal precursor (NT2) cells. J. Neurochem 110, 1828-1841.<br />

Tegenge, M. A., Stern, M., and Bicker, G. (2009). <strong>Nitric</strong> oxide and cyclic nucleotide signal<br />

transduction modulates synaptic vesicle turnover <strong>in</strong> human model neurons. J. Neurochem. 111,<br />

1434-1446.<br />

Tegenge, M.A., Rockel, T.D., Fritsche, E. and Bicker G. <strong>Nitric</strong> oxide signal<strong>in</strong>g as regulator <strong>of</strong><br />

human neuronal progenitor cell migration (In preparation).<br />

Podrygajlo, G., Tegenge, M. A., Gierse, A., Paquet-Durand, F., Tan, S., Bicker, G., and Stern, M.<br />

(2009). Cellular phenotypes <strong>of</strong> human model neurons (NT2) after differentiation <strong>in</strong> aggregate<br />

culture. Cell Tissue Res 336, 439-452.<br />

Abstract presented <strong>in</strong> scientific meet<strong>in</strong>gs:<br />

Tegenge, M.A and Bicker G. (2009). <strong>Nitric</strong> oxide and cyclic guanos<strong>in</strong>e-monophosphate signal<br />

transduction facilitates cell motility and neurite outgrowth <strong>in</strong> differentiat<strong>in</strong>g human model neurons,<br />

Meet<strong>in</strong>g <strong>of</strong> the Society for Neuroscience, Chicago, USA.<br />

Tegenge, M.A and Bicker G. (2009). Pre-synaptic Vesicle Exocytosis <strong>in</strong> <strong>Human</strong> Model Neurons<br />

Generated by Spherical Aggregate Culture Method, 8 th Meet<strong>in</strong>g <strong>of</strong> the German Neuroscience<br />

Society, Gott<strong>in</strong>gen, Germany.<br />

Tegenge, M.A. and Bicker G. (2008). <strong>Nitric</strong> oxide signal transduction facilitates the migration <strong>of</strong><br />

human neuronal precursor (NT2) cells, FENS Abstr. vol.4, 180.3, 6 th FENS Forum <strong>of</strong> European<br />

Neuroscience, Geneva, Switzerland.<br />

Tegenge, M.A. and Bicker G. (2008). <strong>Nitric</strong> oxide signal transduction facilitates the migration <strong>of</strong><br />

human neuronal precursor (NT2) cells, Bra<strong>in</strong>storm<strong>in</strong>g IV, Center for Systems Neuroscience<br />

Hannover, Germany.


Table <strong>of</strong> content<br />

1. Introduction 1<br />

1.1. <strong>Nitric</strong> oxide signal transduction 1<br />

1.1.1. Discovery <strong>of</strong> NO 1<br />

1.1.2. Synthesis <strong>of</strong> NO and activation <strong>of</strong> downstream effectors 1<br />

1.2. NO and early stages <strong>of</strong> CNS development 3<br />

1.2.1. Neuronal precursor proliferation 3<br />

1.2.2. Neuronal migration 4<br />

1.2.3. Neuronal differentiation 5<br />

1.2.4. Synaptogenesis 6<br />

1.3. NO and adult neurogenesis 7<br />

1.4. NO signal<strong>in</strong>g <strong>in</strong> neurodegenerative diseases 8<br />

1.5. <strong>Human</strong> neuronal stem cells as a model <strong>of</strong> develop<strong>in</strong>g nerve cells 9<br />

1.6. Hypothesis and aim <strong>of</strong> the study 11<br />

2. Discussion 12<br />

2.1. <strong>Nitric</strong> oxide and cGMP signal transduction positively regulates<br />

the motility <strong>of</strong> human neuronal precursor (NT2) cells 12<br />

2.2. <strong>Nitric</strong> oxide and cyclic nucleotide signal transduction modulates<br />

synaptic vesicle turnover <strong>in</strong> human model neurons 13<br />

2.3. <strong>Nitric</strong> oxide signal<strong>in</strong>g as regulator <strong>of</strong> human neuronal progenitor cell migration 16<br />

References 18<br />

Acknowledgments 30<br />

Summary 31<br />

Zusammenfassung 33<br />

3. Publications<br />

3.1. <strong>Nitric</strong> oxide and cGMP signal transduction positively regulates the motility <strong>of</strong> human<br />

neuronal precursor (NT2) cells, Tegenge M.A. and Bicker G. (2009), J. Neurochem. 110, 1828-1841.<br />

3.2. <strong>Nitric</strong> oxide and cyclic nucleotide signal transduction modulates synaptic vesicle turnover <strong>in</strong><br />

human model neurons, Tegenge M.A. et al. (2009), J. Neurochem. 111, 1434-1446.<br />

3.3. <strong>Nitric</strong> oxide signal<strong>in</strong>g as regulator <strong>of</strong> human neuronal progenitor cell migration,<br />

Tegenge M.A. et al. (In preparation).


List <strong>of</strong> abbreviations<br />

AD, Alzheimer’s disease<br />

BrdU, 5-bromo-2’-deoxyurid<strong>in</strong>e<br />

cADPR, cyclic adenos<strong>in</strong>e diphosphate ribose<br />

cGMP, cylic guanos<strong>in</strong>e-monophosphate<br />

CICR, Ca (2+)-<strong>in</strong>duced Ca (2+) release<br />

CNS, central nervous system<br />

CNGs, cyclic nucleotide gated ion channels<br />

CREB, cyclic adenos<strong>in</strong>e monophosphate response element b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong><br />

DG, dentate gyrus<br />

DRG, dorsal root ganglion<br />

EC, embryonal carc<strong>in</strong>oma cells<br />

EDRF, endothelial derived relax<strong>in</strong>g factor<br />

eNOS, endothelial nitric oxide synthase<br />

ES, embryonic stem cells<br />

FM1-43, N-(3-triethylammoniumpropyl)-4-(4-(dibutylam<strong>in</strong>o)styryl)pyrid<strong>in</strong>ium dibromide<br />

GTP, guanos<strong>in</strong>e triphosphate<br />

HDAC, histone deacetylase<br />

hNPCs, human neural progenitor cells<br />

iNOS, <strong>in</strong>ducible nitric oxide synthase<br />

IR, immunoreactive<br />

L-NAME, nitro-L-arg<strong>in</strong><strong>in</strong>e methyl ester<br />

MAP, microtubule-associated prote<strong>in</strong><br />

MPTP, 1-methyl-4-phenyl-1,2,3,6- tetrahydropyrid<strong>in</strong>e<br />

7NI, 7-nitro<strong>in</strong>dazole<br />

NMDA, N-methyl D-aspartate<br />

NO, nitric oxide<br />

NOC-18, (2, 2-(Hydroxynitrosohydraz<strong>in</strong>o) bis-ethanam<strong>in</strong>e])<br />

NOS, nitric oxide synthase<br />

nNOS, neuronal nitric oxide synthase<br />

NSC, neuronal stem cells<br />

6-OHDA, 6-hydroxydopam<strong>in</strong>e<br />

PBS, phosphate-buffered sal<strong>in</strong>e<br />

PDEs, cyclic nucleotide phosphodiesterase


PKA, prote<strong>in</strong> k<strong>in</strong>ase A<br />

PKG, prote<strong>in</strong> k<strong>in</strong>ase G<br />

PSD, postsynaptic density<br />

RA, ret<strong>in</strong>oic acid<br />

RyRs, ryanod<strong>in</strong>e receptor<br />

sGC, soluble guanylyl cyclase<br />

SVZ, subventricular zone


1. Introduction<br />

1.1. <strong>Nitric</strong> oxide signal transduction pathway<br />

1.1.1. Discovery <strong>of</strong> NO<br />

Studies <strong>in</strong> 1970s have demonstrated that the smooth muscle-relax<strong>in</strong>g effects <strong>of</strong> nitroglycer<strong>in</strong>e and<br />

other organic nitrate vasodilators <strong>in</strong>volve an active metabolite, nitric oxide (NO), whose properties<br />

are very smilar to a substance named endothelial derived relax<strong>in</strong>g factor (EDRF) (Arnold et al.,<br />

1977; Katsuki et al., 1977). Furchgott and Zawadzki (1980) elegantly showed that relaxation <strong>of</strong><br />

blood vessels by acetylchol<strong>in</strong>e is abolished when the endothelial layer is removed, but can be<br />

rescued by reapply<strong>in</strong>g endothelial cells to the smooth muscle layer. These studies <strong>in</strong>dicated that a<br />

diffusible molecule from the endothelium mediates muscle and blood vessels relaxation, and<br />

attempt to isolate this molecule was proved to be difficult. Subsequent studies directly confirmed<br />

that chemically detectable levels <strong>of</strong> NO released from endothelial cells account for all EDRF<br />

activity (Murad, 1986; Ignarro, 1987; Palmer, 1987). In 1998, a group <strong>of</strong> scientists that <strong>in</strong>clude<br />

Robert Furchgott, Louis Ignarro, and Ferid Murad have been awarded the Nobel Prize <strong>in</strong> Medic<strong>in</strong>e<br />

or Physiology for their discoveries concern<strong>in</strong>g NO as a signal<strong>in</strong>g molecule <strong>in</strong> the cardiovascular<br />

system. This was the first discovery that revealed a simple gas can act as a signal<strong>in</strong>g molecule <strong>in</strong> an<br />

animal, which <strong>in</strong>itiated a remarkable number <strong>of</strong> researches <strong>in</strong> various field.<br />

In the bra<strong>in</strong> Garthwaite et al. (1988) demonstrated that EDRF like diffusible messenger <strong>in</strong>creased<br />

the level <strong>of</strong> cGMP <strong>in</strong> dissociated culture <strong>of</strong> neonatal cerebellar cells. The formation <strong>of</strong> NO from<br />

arg<strong>in</strong><strong>in</strong>e was directly demonstrated <strong>in</strong> cerebellar slices (Bredt and Synder, 1989). Subsequent<br />

studies established NO as unconventional retrograde neurotransmitter that has been implicated to<br />

modulate synaptic plasticity, memory formation, cell proliferation, migration, differentiation and<br />

synaptogenesis (reviewed by Müller, 1997; Enikolopov et al., 1999; Bicker, 2001; 2005; Boehn<strong>in</strong>g<br />

and Snyder, 2003; Cárdenas et al., 2005; Garthwaite, 2008). However, only limited data are<br />

available address<strong>in</strong>g the role <strong>of</strong> NO signal transduction dur<strong>in</strong>g the development <strong>of</strong> the central<br />

nervous system (CNS) <strong>in</strong> both human and experimental animals.<br />

1.1.2. Synthesis <strong>of</strong> NO and activation <strong>of</strong> downstream effectors<br />

NO is synthesized by nitric oxide synthase (NOS), an enzyme that converts L-arg<strong>in</strong><strong>in</strong>e <strong>in</strong>to<br />

equimolar quantities <strong>of</strong> NO and L-citrull<strong>in</strong>e <strong>in</strong> the presence <strong>of</strong> several c<strong>of</strong>actors (Figure 1). There<br />

are three is<strong>of</strong>orms <strong>of</strong> NOS; neuronal NOS (nNOS or NOS I), endothelial NOS (eNOS or NOS III)<br />

and <strong>in</strong>ducible NOS (iNOS or NOS II). Two <strong>of</strong> these enzymes are constitutively expressed <strong>in</strong><br />

1


neuronal and endothelial cells; therefore known as neuronal and endothelial NOS, respectively<br />

(Knowles and Moncada, 1994, Förstermann et al., 1995).<br />

The eNOS is found <strong>in</strong> the caveoli <strong>of</strong> endothelial cells and is activated follow<strong>in</strong>g chol<strong>in</strong>ergic<br />

stimulation and the consequent <strong>in</strong>crease <strong>of</strong> <strong>in</strong>tracellular calcium. The nNOS is found <strong>in</strong> the<br />

cerebellum, the cerebral cortex, sp<strong>in</strong>al cord and <strong>in</strong> various ganglion cells <strong>of</strong> the autonomic nervous<br />

system (Bredt et al., 1990; Grzybicki et al., 1996; Zhou and Zhu, 2009). The nNOS is physically<br />

associated with the N-methyl D-aspartate (NMDA) receptor and postsynaptic density prote<strong>in</strong>-95<br />

(PSD-95) which suggests NMDA activation as a precondition for the synthesis <strong>of</strong> NO (Garthwaite<br />

et al., 1988, Garthwaite, 2008). The <strong>in</strong>ducible NOS (iNOS or NOS II) is formed ma<strong>in</strong>ly <strong>in</strong> immune<br />

cells, such as macrophages and glial cells (Agullo and Garcia, 1992; Simmons and Murphy, 1992).<br />

Unlike eNOS and nNOS, synthesis <strong>of</strong> iNOS mRNA is <strong>in</strong>duced by lipopolysaccharide that activate<br />

its receptors on the surface <strong>of</strong> macrophages and astrocytes (Baltrons et al., 2003, Rettori et al.,<br />

2009).<br />

Although it is difficult to accurately determ<strong>in</strong>e the exact physiological concentrations <strong>of</strong> NO, recent<br />

studies suggested that it may range from 100 pM to 5 nM, orders <strong>of</strong> magnitude lower than<br />

previously thought (Hall and Garthwaite, 2009). Hence, activation <strong>of</strong> its downstream targets<br />

depends on local concentration <strong>of</strong> NO and availability <strong>of</strong> target molecules (Madhusoodanan and<br />

Murad, 2007). The major low concentration physiological target enzyme for NO is the enzyme,<br />

soluble guanylyl cyclase (sGC) (Garthwaite, 2008). sGC is a heterodimeric prote<strong>in</strong> composed <strong>of</strong> α<br />

and β subunits. There are two forms <strong>of</strong> α subunits; the major occurr<strong>in</strong>g α1 and the less abundant α2<br />

which are dimerized to common β subunit. The αβ-heterodimer comprises a haem-b<strong>in</strong>d<strong>in</strong>g region<br />

and a catalytic doma<strong>in</strong> (Haghikia et al., 2007, Garthwaite, 2008). B<strong>in</strong>d<strong>in</strong>g <strong>of</strong> NO to the heme<br />

doma<strong>in</strong> leads to the conversion <strong>of</strong> guanos<strong>in</strong>e triphosphate (GTP) to cylic guanos<strong>in</strong>e-monophosphate<br />

(cGMP) (Figure 1). Genomic deletion <strong>of</strong> the β1 subunit <strong>of</strong> sGC has been implicated to completely<br />

disrupt the NO-cGMP signal<strong>in</strong>g whereas mice lack<strong>in</strong>g both α and β subunits has been employed to<br />

dissect cGMP-<strong>in</strong>dependent action <strong>of</strong> NO (Friebe and Koesl<strong>in</strong>g, 2009). Potential target prote<strong>in</strong>s<br />

downstream <strong>of</strong> cGMP <strong>in</strong>cludes prote<strong>in</strong> k<strong>in</strong>ase G (PKG), cyclic nucleotide gated ion channels<br />

(CNGs) and cyclic nucleotide phosphodiesterase (PDEs). Each <strong>of</strong> these downstream effectors then<br />

transmit the signals to an array <strong>of</strong> <strong>in</strong>tracellular signal<strong>in</strong>g molecules, thereby regulat<strong>in</strong>g<br />

neurotransmission, proliferation, cell migration, differentiation, axon outgrowth and guidance<br />

(Madhusoodanan and Murad , 2007).<br />

In pathological conditions (such as bra<strong>in</strong> ischaemia or neurological disorders) the level <strong>of</strong> NO is<br />

elevated as a result <strong>of</strong> over-activation <strong>of</strong> NMDA receptor <strong>in</strong> neurons or iNOS activation <strong>in</strong> glial<br />

cells. At this high concentration, NO can reacts with superoxide anion to form the very reactive<br />

peroxynitrite that causes neuronal toxicity. NO has also been shown to cause S-nitrosylation <strong>of</strong><br />

2


several prote<strong>in</strong>s that is responsible for neuronal death (Blaise et al., 2005; He et al., 2007; Cho et al.,<br />

2009). Thus, whilst NO plays a physiological role <strong>in</strong> neuronal cell signal<strong>in</strong>g, its over-production<br />

may cause neuronal death.<br />

Figure 1. Schematic draw<strong>in</strong>g <strong>of</strong> the NO signal<strong>in</strong>g cascade. The enzyme nitric oxide synthase (NOS)<br />

is stimulated by calcium–calmodul<strong>in</strong> (Ca/CaM). In the presence <strong>of</strong> O2 and NADPH NO is formed.<br />

NO b<strong>in</strong>ds to the heme-moiety <strong>of</strong> soluble guanylyl cyclase (sGC) result<strong>in</strong>g <strong>in</strong> the stimulation <strong>of</strong> the<br />

enzyme which result <strong>in</strong> elevation <strong>of</strong> cGMP level.<br />

1.2. NO and early stages <strong>of</strong> CNS development<br />

1.2.1. Neuronal precursor proliferation<br />

The complex mammalian CNS, with its diverse cell types and billions <strong>of</strong> connections, undergoes<br />

several cellular stages <strong>of</strong> development. The formation <strong>of</strong> the CNS beg<strong>in</strong>s early <strong>in</strong> development with<br />

the <strong>in</strong>duction <strong>of</strong> the neural ectoderm on the dorsal surface <strong>of</strong> the embryo. Subsequently, the neural<br />

ectoderm plate changes its shape to form a neural groove and eventually, a neural tube. The wall <strong>of</strong><br />

the neural tube is composed <strong>of</strong> germ<strong>in</strong>al cells, collectively called the neuroepithelium, which<br />

produces neurons and glia throughout the CNS (Wessley and De Robertis, 2002; Spitzer, 2006;<br />

Corb<strong>in</strong> et al., 2008). After closure <strong>of</strong> the neural tube at an early embryonic stage, process <strong>of</strong><br />

neuronal proliferation, migration and differentiation occurs sequentially but partially overlapp<strong>in</strong>g<br />

(Spitzer, 2006; Ayala et al., 2007).<br />

Dur<strong>in</strong>g development, transient expression <strong>of</strong> nNOS has been demonstrated <strong>in</strong> different bra<strong>in</strong> areas,<br />

which implicate functional role <strong>of</strong> NO <strong>in</strong> embryonic neural tissue formation (Bredt and Snyder,<br />

3


1994; Santacana et al., 1998; Chen et al., 2004; Nott and Riccio, 2009). In l<strong>in</strong>e with this, NO has<br />

been shown to suppress the proliferation <strong>of</strong> neural precursor cells <strong>in</strong> develop<strong>in</strong>g Drosophila<br />

imag<strong>in</strong>al disks (Kuz<strong>in</strong> et al., 1996; Enikolopov et al., 1999), Xenopus tadpole optic tectum<br />

(Peunova et al., 2001, Peunova et al., 2007) and cerebellar granule cells <strong>of</strong> new born rats (Ciani et<br />

al., 2006). In develop<strong>in</strong>g chicken neural tube NO is endogenously produced which has been<br />

proposed to regulate cell-cycle progression. It has been demonstrated that high NO levels promoted<br />

entry <strong>in</strong>to S phase basally, whereas low levels <strong>of</strong> NO facilitated entry <strong>in</strong>to mitosis apically (Traister<br />

et al., 2002). Moreover, <strong>in</strong> cultured human neuroblastoma cell l<strong>in</strong>es NO has been showed to <strong>in</strong>hibit<br />

neuronal precursor cell proliferation (Murillo-Carretero et al, 2002; Ciani et al., 2004). Treatment<br />

<strong>of</strong> dissociated mouse cortical neuroepithelial cluster cell cultures and rat cerebellar granule cells<br />

with NOS <strong>in</strong>hibitor or scaveng<strong>in</strong>g <strong>of</strong> NO resulted <strong>in</strong> enhanced proliferation (Cheng et al., 2003;<br />

Ciani et al., 2004). These studies show the anti-proliferative role <strong>of</strong> endogenous NO <strong>in</strong> neuronal<br />

precursor cells; however, the mechanism is not fully understood. In develop<strong>in</strong>g Xenopus, cerebellar<br />

granule cells and human neuroblastoma cells, the cGMP/PKG pathway has been suggested to<br />

mediate the anti-proliferative role <strong>of</strong> NO (Peunova et al., 2007; Ciani et al., 2004; Ciani et al.,<br />

2006). In both cerebellar cells and neuroblastoma culture the oncogenic transcription factor, N-<br />

Myc, has been shown to be over expressed due to down regulation <strong>of</strong> cGMP/PKG pathway<br />

(Contestabile A., 2008). On the contrary, other studies demonstrated that the anti-proliferative effect<br />

<strong>of</strong> NO is cGMP <strong>in</strong>dependent (Phung et al., 1999; Young et al., 2000; Gibbs, 2003).<br />

1.2.2. Neuronal migration<br />

Neurons and neuronal precursor cell migrate long distances along the dorsal-ventral and anterior-<br />

posterior axes <strong>of</strong> the nervous system dur<strong>in</strong>g the early embryonic period, which is important for the<br />

formation <strong>of</strong> functional neuronal network. Two modes <strong>of</strong> neuronal migration has been identified:<br />

radial migration, <strong>in</strong> which cells migrate from the progenitor zone toward the surface <strong>of</strong> the bra<strong>in</strong><br />

follow<strong>in</strong>g the radial layout <strong>of</strong> the neural tube; and tangential migration, <strong>in</strong> which cells migrate<br />

orthogonally to the direction <strong>of</strong> radial migration (Hatten, 1999; Mar<strong>in</strong> and Rubenste<strong>in</strong>, 2003; Ayala<br />

et al., 2007; Met<strong>in</strong> et al., 2008). In the develop<strong>in</strong>g rat bra<strong>in</strong>, nNOS reaches the highest expression<br />

level between embryonic day 16 and postnatal day 0, and this expression corresponds with the<br />

migration <strong>of</strong> neuronal precursors from the ventricular zone to the external layers <strong>of</strong> the cortex<br />

(Bredt and Snyder, 1994; Nott et al., 2008).<br />

The first functional evidence on the role <strong>of</strong> NO <strong>in</strong> neuronal migration came from study conducted<br />

on migrat<strong>in</strong>g granule cells <strong>of</strong> rat (Tanaka et al., 1994). The migration <strong>of</strong> cerebellar granule cells was<br />

decreased <strong>in</strong> the presence <strong>of</strong> NOS <strong>in</strong>hibitor or haemoglob<strong>in</strong> that scavenges NO (Tanaka et al.,<br />

4


1994). In embryonic grasshopper NO has been demonstrated to be essential <strong>in</strong> enteric neuron<br />

migration (Haase and Bicker, 2003; Knipp and Bicker, 2009). They showed that pharmacological<br />

<strong>in</strong>hibition <strong>of</strong> NOS or sGC blocks neuronal migration, whereas the block<strong>in</strong>g can be rescued by<br />

apply<strong>in</strong>g sGC activators or cGMP analogue suggest<strong>in</strong>g positive regulatory role <strong>of</strong> NO/cGMP<br />

signal<strong>in</strong>g <strong>in</strong> neuronal migration. In develop<strong>in</strong>g Xenopus, NO has been shown to facilitate neuronal<br />

cell migration (Peunova et al., 2007). Recent studies us<strong>in</strong>g genetic approaches <strong>in</strong>dicated that nNOS<br />

is essential for cortical development (Nott et al., 2008; 2009). Here, radial migration <strong>of</strong> cortical<br />

progenitors was impaired <strong>in</strong> nNOS null mice. Moreover, ex vivo electroporation <strong>of</strong> the mutant, NO<strong>in</strong>sensitive<br />

histone deacetylase 2 (HDAC2) <strong>in</strong> cortical progenitors impaired radial migration<br />

<strong>in</strong>dicat<strong>in</strong>g that NO positively regulates neuronal migration via S-nitrosylation (Nott et al., 2009). In<br />

the fetal human sp<strong>in</strong>al cord, some neurons expressed NOS as they migrate to their f<strong>in</strong>al dest<strong>in</strong>ation,<br />

which suggest that NO may <strong>in</strong>volve <strong>in</strong> neuronal migration dur<strong>in</strong>g human nerve cell development<br />

(Foster and Phelps, 2000).<br />

1.2.3. Neuronal differentiation<br />

Neuronal differentiation has many components, which <strong>in</strong>cludes maturation <strong>of</strong> electrical excitability,<br />

neurotransmitter specification, outgrowth <strong>of</strong> axon and dendrites (Spitzer, 2006). Evidence<br />

support<strong>in</strong>g that NO mediate neuronal differentiation came from nerve growth factor (NGF) <strong>in</strong>duced<br />

differentiat<strong>in</strong>g PC12 cells (Pennova and Enikolopov, 1995). In these cells the expression <strong>of</strong> the<br />

three is<strong>of</strong>orm <strong>of</strong> NOS corresponds with NGF <strong>in</strong>duced neuronal differentiation. Moreover,<br />

application <strong>of</strong> NOS <strong>in</strong>hibitor <strong>in</strong> PC12 culture <strong>in</strong>creased cell proliferation but <strong>in</strong>hibit differentiation<br />

<strong>in</strong>to neurons. Subsequent studies <strong>in</strong>dicated that NO <strong>in</strong>duce differentiation <strong>in</strong>to neuronal phenotypes<br />

while <strong>in</strong>hibit<strong>in</strong>g cell proliferation <strong>in</strong> various models (Ghigo et al., 1998; Phung et al., 1999; Cheng,<br />

2003; Ciani et al., 2004; Ciani et al., 2006, Evangelopoulos et al., 2010). The <strong>in</strong>duction <strong>of</strong> NOS<br />

expression <strong>in</strong> response to NGF has been shown to be mediated through a comb<strong>in</strong>ation <strong>of</strong> the<br />

neurotrophic tyros<strong>in</strong>e k<strong>in</strong>ase receptor type 1 (TrkA) and Ras-Erk members <strong>of</strong> the mitogen activated<br />

prote<strong>in</strong> k<strong>in</strong>ase (MAP k<strong>in</strong>ase) signal<strong>in</strong>g cascade (Bulseco et al., 2001; Schonh<strong>of</strong>f et al., 2001).<br />

Neuronal growth cones, specialized fan-shaped structures at the tips <strong>of</strong> grow<strong>in</strong>g neurites, are<br />

important for axon elongation and guidance. Filopodia, rod-like projections from growth cones,<br />

responds to various extracellular cues dur<strong>in</strong>g neurite elogation and guidance (Goldberg and<br />

Burmeister, 1989; Kater and Rehder, 1995). Dur<strong>in</strong>g the early development <strong>of</strong> primary<br />

somatosensory neurons, nNOS is expressed at time <strong>of</strong> neurite extension (Thippeswamy and Morris,<br />

2002). Several <strong>in</strong> vitro studies demonstrated that growth cones <strong>of</strong> rat dorsal root ganglion (DRG),<br />

Xenopus and chick ret<strong>in</strong>al ganglion and Helisoma buccal ganglion collapsed <strong>in</strong> the presence <strong>of</strong> high<br />

5


concentration <strong>of</strong> NO (Hess et al., 1993; Renteria and Constant<strong>in</strong>e-Paton, 1996; Ernst et al., 2000;<br />

He et al., 2002; Trimm and Rehder, 2004). Recently, S-nitrosylation <strong>of</strong> microtubule-associated<br />

prote<strong>in</strong> 1B (MAP1B) has been suggested to mediate NO-<strong>in</strong>duced axon retraction <strong>in</strong> cultured<br />

vertebrate neurons (Stroissnigg et al., 2007). In Helisoma buccal ganglion, NO regulates growth<br />

cone filopodial behavior via sGC, PKG and cyclic adenos<strong>in</strong>e diphosphate ribose (cADPR), which<br />

causes the release <strong>of</strong> calcium from <strong>in</strong>tracellular stores via the ryanod<strong>in</strong>e receptor (RyRs)<br />

(Welshhans and Rehder, 2005). The downstream effector prote<strong>in</strong>s for NO/cGMP signal<strong>in</strong>g, PKG<br />

(Yue et al., 2008) and CNGs (Togashi et al., 2008) have been shown to mediate ephr<strong>in</strong>-A5-<strong>in</strong>duced<br />

growth cone collapse and Sema3A-<strong>in</strong>duced growth cone repulsion, respectively. Furthermore, the<br />

NO/cGMP signal<strong>in</strong>g has been shown to negatively regulate the Ca 2+ -<strong>in</strong>duced Ca 2+ release (CICR)<br />

through RyRs to control directional polarity <strong>of</strong> DRG axon guidance (Tojima et al., 2009). Here, a<br />

Ca 2+ signal produced by photolys<strong>in</strong>g caged Ca 2+ caused growth cone repulsion on lam<strong>in</strong><strong>in</strong> substrate<br />

which was converted <strong>in</strong>to attraction by pharmacological block<strong>in</strong>g <strong>of</strong> NO/cGMP pathway or genetic<br />

deletion <strong>of</strong> nNOS.<br />

On the other hand, studies performed <strong>in</strong> vitro on PC12 cells (H<strong>in</strong>dley et al., 1997; Rialas et al, 2000;<br />

Yamazaki et al., 2001) and neuroblastoma cells (Evangelopoulos et al., 2010) <strong>in</strong>dicated that NO<br />

<strong>in</strong>creases neurite outgrowth. In develop<strong>in</strong>g antenna <strong>of</strong> the grasshopper embryo where two sibl<strong>in</strong>gs<br />

<strong>of</strong> pioneer neurons establish the first two axonal pathways to the CNS, NO/cGMP signal<strong>in</strong>g has<br />

been shown to mediate axonogenesis (Seidel and Bicker, 2000). Here, pharmacological <strong>in</strong>hibition<br />

<strong>of</strong> NOS and sGC resulted <strong>in</strong> abnormal pattern <strong>of</strong> pathf<strong>in</strong>d<strong>in</strong>g, loss <strong>of</strong> axon emergence and axon<br />

retraction suggest<strong>in</strong>g that NO/cGMP signal<strong>in</strong>g is a positive regulator <strong>of</strong> neurite elongation. Recent<br />

study <strong>in</strong>dicated that S-nitrosylation HDAC2 caused by neurotroph<strong>in</strong> <strong>in</strong>duced NO signal<strong>in</strong>g is<br />

necessary for dendritic outgrowth <strong>of</strong> embryonic cortical neurons (Nott et al., 2008). In models <strong>of</strong><br />

CNS <strong>in</strong>jury, the regeneration <strong>of</strong> axons <strong>in</strong> embryonic <strong>in</strong>sect (Stern and Bicker, 2008) and optic nerve<br />

<strong>in</strong> the goldfish (Koriyama et al., 2009) was facilitated by NO/cGMP signal<strong>in</strong>g pathway.<br />

1.2.4. Synaptogenesis<br />

Synaptogenesis can be def<strong>in</strong>ed as the assembly <strong>of</strong> pre- and postsynaptic prote<strong>in</strong>s <strong>in</strong>to the highly<br />

specific structure <strong>of</strong> the synapse. These major components <strong>of</strong> glutamatergic synapses <strong>in</strong>cludes;<br />

synaptic vesicles (SVs), glutamate receptors, active zone prote<strong>in</strong>s, postsynaptic density (PSD)<br />

scaffold<strong>in</strong>g prote<strong>in</strong>s, and trans-synaptic adhesion molecules. The pre-and postsynaptic components<br />

<strong>of</strong> a synapse must accumulate at sites <strong>of</strong> physical contact between axons and dendrites with precise<br />

tim<strong>in</strong>g (McAllister, 2007). Dur<strong>in</strong>g development, synaptogenesis is tightly coupled to neuronal<br />

differentiation and formation <strong>of</strong> neuronal circuitry. For example, shortly after neurons differentiate<br />

6


and extend axonal and dendritic processes, many <strong>of</strong> the genes encod<strong>in</strong>g synaptic prote<strong>in</strong>s are turned<br />

on, result<strong>in</strong>g <strong>in</strong> the formation, accumulation, and directional traffick<strong>in</strong>g <strong>of</strong> vesicles carry<strong>in</strong>g pre- and<br />

postsynaptic prote<strong>in</strong> complexes. Dur<strong>in</strong>g this time, the specification <strong>of</strong> correct neuronal connections<br />

is determ<strong>in</strong>ed, as axons and dendrites make contact and establish <strong>in</strong>itial, <strong>of</strong>ten transient, synapses.<br />

Thus, the process <strong>of</strong> synaptogenesis <strong>in</strong>volves a myriad <strong>of</strong> secreted factors, receptors, and signal<strong>in</strong>g<br />

molecules that make neurons receptive to form synapses (Waites et al., 2005).<br />

Transient expression <strong>of</strong> NOS has been suggested to correspond with the period <strong>of</strong> active synapse<br />

formation (Williams et al., 1994; Ogilvie et al., 1995; Sporns and Jenk<strong>in</strong>son, 1997; Gibbs and<br />

Truman, 1998; Bicker, 2005). Moreover, <strong>in</strong> embryonic grasshoppers, synaptogenesis correlates with<br />

a phase when many identifiable nerve cell types respond to NO by produc<strong>in</strong>g cGMP (Truman et al.,<br />

1996; Ball and Truman, 1998). The dynamic regulation <strong>of</strong> NO-<strong>in</strong>duced cGMP formation dur<strong>in</strong>g<br />

synaptic maturation has been also described <strong>in</strong> the embryonic grasshopper bra<strong>in</strong> and at the larval<br />

neuromuscular junction <strong>of</strong> Drosophila (Wildemann and Bicker, 1999; Seidel and Bicker, 2002). NO<br />

signal<strong>in</strong>g has been implicated to play an important role <strong>in</strong> activity-dependent synaptic maturation<br />

(Nikonenko et al., 2003). Here, <strong>in</strong>hibition <strong>of</strong> NOS has been demonstrated to <strong>in</strong>terfere with calciumdependent<br />

growth <strong>of</strong> filopodia-like outgrowth and presynaptic remodell<strong>in</strong>g <strong>of</strong> varicosites while NO<br />

donors could facilitate them.<br />

Excitatory synapse formation, elim<strong>in</strong>ation and remodell<strong>in</strong>g on dendritic sp<strong>in</strong>es represent a<br />

cont<strong>in</strong>uous process, which shapes the organization <strong>of</strong> synaptic networks dur<strong>in</strong>g development<br />

(Yoshihara et al., 2009). NO is produced at excitatory synapses by nNOS which is localized to the<br />

synapse through its <strong>in</strong>teraction with the second PDZ doma<strong>in</strong> <strong>of</strong> PSD-95. Overexpression <strong>of</strong> PSD-95<br />

resulted <strong>in</strong> the rise <strong>of</strong> nNOS, lead<strong>in</strong>g to the formation <strong>of</strong> sp<strong>in</strong>es which became <strong>in</strong>nervated by<br />

multiple presynaptic partners (Nikonenko et al., 2008). A similar observation was made when the<br />

concentration <strong>of</strong> NO is <strong>in</strong>creased <strong>in</strong> the tissue by exogenous application <strong>of</strong> a NO donor. Conversely,<br />

blockade <strong>of</strong> nNOS <strong>in</strong>terferes with synapse formation and resulted <strong>in</strong> a loss <strong>of</strong> sp<strong>in</strong>es. NO produced<br />

by expression <strong>of</strong> the PSD <strong>in</strong> the newly formed sp<strong>in</strong>e could thus represent a retrograde signal<br />

stimulat<strong>in</strong>g nearby axons to differentiate and form a presynaptic term<strong>in</strong>al (Nikonenko et al., 2008;<br />

Yoshihara et al., 2009).<br />

1.3. NO and adult neurogenesis<br />

The dentate gyrus (DG) <strong>of</strong> the hippocampus and the subventricular zone (SVZ) possess the capacity<br />

to generate new neurons <strong>in</strong> adult animals (Gage, 2000; Curtis et al., 2003). The whole proliferation<br />

area <strong>in</strong> both SVZ and DG has been suggested to be under the <strong>in</strong>fluence <strong>of</strong> NO (Estrada and Murillo-<br />

Carretero, 2005). Neuroanatomical studies <strong>in</strong>dicate that nNOS and sGC is expressed <strong>in</strong> adult SVZ<br />

(Moreno-López et al., 2000; Gutierrez-Mec<strong>in</strong>as et al., 2007). The exist<strong>in</strong>g evidence from several<br />

7


laboratories implicate that NO exerts a negative control on proliferation <strong>of</strong> cells <strong>in</strong> SVZ (Packer et<br />

al., 2003; Cheng et al., 2003; Moreno-Lopez et al., 2004).<br />

Reports on the role <strong>of</strong> NO <strong>in</strong> adult hippocampal neurogenesis; however, are contradictory. Systemic<br />

application <strong>of</strong> NOS <strong>in</strong>hibitors, nitro-L-arg<strong>in</strong><strong>in</strong>e methyl ester (L-NAME) or 7- nitro<strong>in</strong>dazole (7NI) to<br />

adult rodents did not alter the number <strong>of</strong> mitotic cells <strong>in</strong> the DG <strong>of</strong> the hippocampus (Moreno-<br />

Lopez et al., 2004) but the density <strong>of</strong> Bromo-deoxyurid<strong>in</strong>e (BrdU) positive cells <strong>in</strong>creased<br />

significantly after chronic <strong>in</strong>hibition <strong>of</strong> NOS ( Park et al., 2003). nNOS derived NO has been shown<br />

to <strong>in</strong>hibit adult neurogenesis <strong>in</strong> DG by down-regulat<strong>in</strong>g cyclic AMP response element b<strong>in</strong>d<strong>in</strong>g<br />

prote<strong>in</strong> (CREB ) phosphorylation (Zhu, 2006). The use <strong>of</strong> mutant mice lack<strong>in</strong>g functional nNOS<br />

resulted <strong>in</strong> <strong>in</strong>creased proliferation <strong>in</strong> DG (Packer et al., 2003) whereas mice lack<strong>in</strong>g eNOS showed a<br />

significant reduction <strong>in</strong> neuronal progenitor cell proliferation <strong>in</strong> the same structure (Reif et al.,<br />

2004). NSC proliferation and survival <strong>of</strong> new born neurons <strong>in</strong> the hippocampus were <strong>in</strong>vestigated <strong>in</strong><br />

nNOS knock out mice and double knock out mice (nNOS/eNOS). The proliferation <strong>of</strong> NSC was not<br />

significantly altered <strong>in</strong> nNOS knock out mice but survival <strong>of</strong> newly formed neurons was<br />

substantially higher (Fritzen et al., 2007). In contrast, nNOS/eNOS double knockout mice had<br />

significantly decreased survival rates.<br />

Bra<strong>in</strong> <strong>in</strong>juries such as ischemia, traumas and seizures are characterized by overproduction <strong>of</strong> NO.<br />

This could be as a result <strong>of</strong> nNOS activation follow<strong>in</strong>g massive release <strong>of</strong> glutamate, up-regulation<br />

<strong>of</strong> nNOS and iNOS (Estrada and Murillo-Carretero, 2005). It has been demonstrated that expression<br />

<strong>of</strong> iNOS is necessary for ischemia-stimulated cell birth <strong>in</strong> the DG after focal cerebral ischemia (Zhu<br />

et al., 2003; Luo et al., 2005). Similarly, exogenous NO source reported to <strong>in</strong>crease cell<br />

proliferation and neuroblast migration <strong>in</strong> SVZ and DG (Zhang et al., 2001; Chen et al., 2004; Cui et<br />

al., 2009). In contrast, decreased nNOS expression was observed <strong>in</strong> SVZ which corresponds to<br />

<strong>in</strong>creased cell proliferation but reduced cell migration after focal cerebral ischemia <strong>in</strong> rats (Sun et<br />

al., 2005; Zhang et al., 2007). A recent study demonstrated that cerebral ischemia reduces<br />

hippocampal nNOS expression, and <strong>in</strong>hibition <strong>of</strong> nNOS ameliorates ischemic <strong>in</strong>jury, stimulates cell<br />

proliferation, and up-regulates iNOS expression and CREB phosphorylation (Luo et al., 2007).<br />

Thus, it appears that the source <strong>of</strong> NO (whether it is from eNOS, iNOS or nNOS) and condition <strong>of</strong><br />

the animal (health versus disease) dictate the role <strong>of</strong> NO <strong>in</strong> adult neurogenesis.<br />

1.4. NO signal<strong>in</strong>g <strong>in</strong> neurodegenerative diseases<br />

S<strong>in</strong>ce NO is an important regulator <strong>of</strong> nervous system functions, substantial changes <strong>in</strong> NO and<br />

cGMP synthesis may lead to nervous system degeneration. Several studies have <strong>in</strong>dicated that NO<br />

may directly or <strong>in</strong>directly play a key role <strong>in</strong> the pathogenesis <strong>of</strong> a number <strong>of</strong> neurodegenerative<br />

8


diseases <strong>in</strong>clud<strong>in</strong>g Alzheimers, Amyotrophic lateral sclerosis (ALS) and Park<strong>in</strong>son diseases<br />

(Madhusoodanan and Murad, 2007).<br />

The role <strong>of</strong> NO <strong>in</strong> the development <strong>of</strong> Park<strong>in</strong>son disease has been extensively studied <strong>in</strong> 1-methyl-<br />

4-phenyl-1,2,3,6- tetrahydropyrid<strong>in</strong>e (MPTP) and 6-hydroxydopam<strong>in</strong>e (6-OHDA) <strong>in</strong>duced animal<br />

models. Previous studies implicated that mice lack<strong>in</strong>g nNOS are more resistant to MPTP <strong>in</strong>duced<br />

neurotoxicity (Przedborski et al., 1996; Matthews et al., 1997). The nNOS <strong>in</strong>hibitor 7NI has been<br />

shown to protect aga<strong>in</strong>st MPTP-<strong>in</strong>duced neurotoxicity <strong>in</strong> experimental animals (Kurosaki et al.,<br />

2002; Watanabe et al., 2008). 7NI has been also showed to protect from 6-OHDA <strong>in</strong>duced<br />

neurodegeneration whereas NO donor worsened the neurodegeneration (Di Matteo et al., 2009).<br />

The majority <strong>of</strong> studies <strong>in</strong>dicated cGMP <strong>in</strong>dependent mechanism <strong>of</strong> action <strong>of</strong> NO <strong>in</strong> mediat<strong>in</strong>g<br />

MPTP <strong>in</strong>duced neurotoxicity. These cGMP-<strong>in</strong>dependent actions <strong>of</strong> NO <strong>in</strong>clude oxidation <strong>of</strong> thiols,<br />

S-nitrosylation and nitration <strong>of</strong> prote<strong>in</strong>s (Madhusoodanan and Murad, 2007).<br />

In Alzheimer’s disease (AD) all the three is<strong>of</strong>orms <strong>of</strong> NOS have been shown to be elevated,<br />

<strong>in</strong>dicat<strong>in</strong>g an important role for NO <strong>in</strong> the pathomechanism <strong>of</strong> AD (Thorns et al., 1998; Lüth 2001).<br />

Nitrotyros<strong>in</strong>e which has been found <strong>in</strong> AD patients was highly co-localized with nNOS <strong>in</strong> cortical<br />

pyramidal cells (Simic et al., 2000; Sultana et al., 2006). Immunocytochemical methods<br />

demonstrated that the nNOS, iNOS and nitrotyros<strong>in</strong>e immunopositive cell bodies over the entire<br />

chronic AD patients bra<strong>in</strong>s (Fernández-Vizarra et al., 2004).<br />

S<strong>in</strong>ce the expression and activity <strong>of</strong> nNOS is <strong>in</strong>creased <strong>in</strong> many CNS diseases, the NO/cGMP<br />

signal<strong>in</strong>g pathway could be a putative target for develop<strong>in</strong>g drugs. In l<strong>in</strong>e with this, several nNOS<br />

<strong>in</strong>hibitors were developed but appear to have side effects (Zhou and Zhu, 2009). To m<strong>in</strong>imize<br />

unwanted effects <strong>in</strong> recent years more attention is given toward design<strong>in</strong>g a selective nNOS<br />

<strong>in</strong>hibitor with a potential implication to target neurodegenerative diseases (Silverman, 2009). On<br />

the other hand NO synthesized by eNOS has been shown to be neuroprotective <strong>in</strong> an <strong>in</strong> vitro model<br />

<strong>of</strong> ALS (De Palma et al., 2008). Here, overexpression <strong>of</strong> eNOS by neurons has been suggested as a<br />

neuroprotective mechanism <strong>in</strong> neurodegenerative diseases. Changes <strong>in</strong> histone acetylation that<br />

certa<strong>in</strong>ly <strong>in</strong>volve NO signal<strong>in</strong>g have been implicated to partially improve memory loss and<br />

neurodegeneration (Fischer et al., 2007, Riccio et al., 2006, Nott et al., 2008). Thus, careful<br />

manipulation <strong>of</strong> NO signal<strong>in</strong>g may prove to be a potential therapeutic approach <strong>in</strong><br />

neurodegenerative disorders.<br />

1.5. <strong>Human</strong> neuronal stem cells as a model <strong>of</strong> develop<strong>in</strong>g nerve cells<br />

Stem cells are def<strong>in</strong>ed as precursor cells that have the ability to self-renew and to generate various<br />

differentiat<strong>in</strong>g cell types. Pluripotent stem cells can give rise theoretically to every cell type <strong>in</strong> the<br />

animal body <strong>in</strong>clud<strong>in</strong>g neurons. Three types <strong>of</strong> mammalian pluripotent stem cell l<strong>in</strong>es have been<br />

9


isolated: embryonal carc<strong>in</strong>oma (EC) cells, the stem cells <strong>of</strong> testicular tumours; embryonic stem (ES)<br />

cells derived from pre-implantation embryos; and embryonic germ (EG) cells derived from<br />

primordial germ cells (PGCs) <strong>of</strong> the post-implantation embryo (Donovan and Gearhart, 2001).<br />

Neuronal stem cells (NSC) are pluripotent cells that have the ability to generate the three major cell<br />

types <strong>of</strong> the CNS; neurons, astrocytes and oligodendrocytes. The search for a therapeutic option to<br />

treat neurodegenerative diseases caused an expansion <strong>of</strong> NSC biology, which revealed numerous<br />

transcription factors, growth factors and signal<strong>in</strong>g pathways <strong>in</strong>volved <strong>in</strong> development and<br />

regeneration <strong>of</strong> CNS (Gage, 2000; Curtis et al., 2003; J<strong>in</strong> et al., 2004; Shan et al., 2006; Boucherie<br />

and Hermans, 2009).<br />

Neuronal stem cells could be employed as important tool to dissect the development <strong>of</strong> human<br />

nervous system. For example, several <strong>of</strong> the exist<strong>in</strong>g human EC stem cell l<strong>in</strong>es provide robust and<br />

simple culture systems to study certa<strong>in</strong> aspects <strong>of</strong> cellular differentiation that mimic vertebrate<br />

neurogenesis (Przyborski et al., 2004). The teratocarc<strong>in</strong>oma cell l<strong>in</strong>e Ntera2 (NT2) which is<br />

derived from a human testicular cancer can be <strong>in</strong>duced to differentiate <strong>in</strong>to fully functional<br />

postmitotic neurons and other cell types <strong>of</strong> the neuronal l<strong>in</strong>eages (Andrews 1984; Pleasure et al.,<br />

1992; Paquet-Durand and Bicker, 2007). NT2 cells shares many similarity with human embryonic<br />

stem cells (Schwartz et al., 2005) and differentiation <strong>of</strong> NT2 cells <strong>in</strong>to neurons has been suggested<br />

to resemble vertebrate neurogenesis (Przyborski et al., 2000; Houldsworth et al., 2002; Przyborski<br />

et al. 2003). Moreover, fully differentiated NT2 neurons have been shown to express a variety <strong>of</strong><br />

neurotransmitters <strong>in</strong> vitro (Guillema<strong>in</strong> et al., 2000; Podrygajlo et al., 2009) and form functional<br />

synapse (Hartley et al., 1998; Podrygajlo et al., 2010).<br />

More restricted stem cells with developmental potential can also be obta<strong>in</strong>ed from a variety <strong>of</strong><br />

tissues at different stages <strong>of</strong> development, <strong>in</strong>clud<strong>in</strong>g mature tissues from adult and fetal organisms.<br />

For example, fetal human neural progenitor cells (hNPCs) that have the capacity to give neurons,<br />

astrocytes and oligodendrocytes have been shown to mimic early develop<strong>in</strong>g human bra<strong>in</strong> with<br />

respect to cell proliferation, migration and differentiation (Fritsche et al., 2005; Moors et al., 2007;<br />

Moors et al., 2009). NSCs also exist with<strong>in</strong> limited regions <strong>of</strong> the adult CNS (SVZ and<br />

hippocampus), and it is possible to isolate and expand these cells to give rise to progenitor cells<br />

restricted to def<strong>in</strong>ed neural l<strong>in</strong>eages, neuronal and glial cells (Gage, 2000).<br />

Another promis<strong>in</strong>g model for develop<strong>in</strong>g human neurons is <strong>in</strong>duced pluripotent stem cells (iPS)<br />

obta<strong>in</strong>ed by molecular reprogramm<strong>in</strong>g <strong>of</strong> somatic cells. These cells have been successfully<br />

differentiated <strong>in</strong>to dopam<strong>in</strong>ergic neurons and functionally <strong>in</strong>tegrated <strong>in</strong>to rat bra<strong>in</strong> (Wern<strong>in</strong>g et al.,<br />

2008).<br />

10


1.6. Hypothesis and aim <strong>of</strong> the study<br />

The enzymes that synthesize NO and its target enzyme soluble guanylyl cyclase have been shown<br />

to be expressed dur<strong>in</strong>g development <strong>of</strong> nervous system <strong>in</strong> both vertebrate and <strong>in</strong>vertebrate.<br />

Functional studies that <strong>in</strong>clude neuronal proliferation, migration and synaptogenesis strongly<br />

suggest that NO signal<strong>in</strong>g is <strong>in</strong>volved <strong>in</strong> the early development <strong>of</strong> the nervous system. However, it<br />

is unknown whether NO signal<strong>in</strong>g plays a functional role <strong>in</strong> early develop<strong>in</strong>g human nerve cells.<br />

S<strong>in</strong>ce most <strong>of</strong> the molecules <strong>in</strong>volved <strong>in</strong> neuronal migration, axon guidance and synaptogenesis are<br />

conserved and some neurons <strong>in</strong> fetal human sp<strong>in</strong>al cord express NOS as they migrate to their f<strong>in</strong>al<br />

dest<strong>in</strong>ation, we have hypothesized that NO may play functional role to modulate neuronal precursor<br />

migration and synaptic maturation <strong>in</strong> develop<strong>in</strong>g human bra<strong>in</strong>. To test the hypothesis I have used<br />

two models <strong>of</strong> stem cell derived human neuronal culture. The NT2 spherical aggregates and fully<br />

differentiated NT2 neurons which are obta<strong>in</strong>ed from human teratocarc<strong>in</strong>oma cell l<strong>in</strong>e (Ntera2).<br />

Additionally, I have used fetal human neural progenitor cells (hNPCs) which was ma<strong>in</strong>ta<strong>in</strong>ed as<br />

free-float<strong>in</strong>g three-dimensional neurosphere primary culture.<br />

The aim <strong>of</strong> this work was to test whether the gaseous messenger NO modulates the proliferation,<br />

migration, differentiation and pre-synaptic release <strong>in</strong> develop<strong>in</strong>g human model neurons. Specifically<br />

we tested the follow<strong>in</strong>g hypothesis:<br />

1. The motility <strong>of</strong> cells from human neuronal precursor (NT2) cells spherical aggregate is<br />

regulated by NO and cGMP signal transduction dur<strong>in</strong>g differentiation <strong>in</strong>to neuronal<br />

phenotypes.<br />

2. The gaseous messenger NO and cyclic nucleotide signal<strong>in</strong>g <strong>in</strong>duces pre-synaptic vesicle<br />

release <strong>in</strong> develop<strong>in</strong>g human model neurons.<br />

3. The migration <strong>of</strong> cells from fetal human neural progenitor cells is regulated by NO<br />

signal<strong>in</strong>g.<br />

11


2. Discussion<br />

2.1. <strong>Nitric</strong> oxide and cGMP signal transduction positively regulates the motility <strong>of</strong> human<br />

neuronal precursor (NT2) cells<br />

Neuronal migration is a critical event dur<strong>in</strong>g early development <strong>of</strong> the nervous system. <strong>Human</strong><br />

neurological disorders such as lissencephaly, cortical heterotopias and microcephaly are associated<br />

with defective neuronal migration (Ayala R et al., 2007; Mèt<strong>in</strong> C., et al., 2009). Dur<strong>in</strong>g<br />

development, neurons and neuronal precursor cells migrate long distances <strong>in</strong> response to specific<br />

external signals. This coord<strong>in</strong>ated movement is subjected to modulation by <strong>in</strong>tra-and <strong>in</strong>tercellular<br />

signals. The gaseous messenger NO has been implicated to regulate neuronal migration <strong>in</strong> both<br />

vertebrates and <strong>in</strong>vertebrates (Tanaka et al., 1994; Haase and Bicker, 2003; Peunova et al., 2007;<br />

Knipp and Bicker, 2009). In this study we showed for the first time that the NO/cGMP/PKG<br />

signal<strong>in</strong>g pathway facilitates the migration <strong>of</strong> human neuronal precursor (NT2) cells. The NT2 cells<br />

which were cultured as three-dimensional free-float<strong>in</strong>g spherical aggregate exhibit many feature <strong>of</strong><br />

develop<strong>in</strong>g bra<strong>in</strong> cells. Firstly, NT2 cell spheres express both precursor (nest<strong>in</strong>) and early neuronal<br />

marker (�III-tubul<strong>in</strong>) and <strong>in</strong>corporate BrdU. Upon cultur<strong>in</strong>g on adhesive substrates such as<br />

Matrigel, cells migrate out <strong>of</strong> the NT2 cell spheres, elaborate neuronal processes and differentiate<br />

<strong>in</strong>to fully functional neurons (Tegenge and Bicker, 2009; Tegenge et al., 2009).<br />

Subpopulations <strong>of</strong> the migrated cells and cells with<strong>in</strong> the NT2 aggregate express NO-sensitive sGC<br />

that synthesize <strong>in</strong>creas<strong>in</strong>g level <strong>of</strong> cGMP (Fig.4, Tegenge and Bicker, 2009). Application <strong>of</strong> enzyme<br />

<strong>in</strong>hibitors <strong>of</strong> nNOS, sGC and PKG blocked the migration <strong>of</strong> cells out <strong>of</strong> NT2 aggregates. Whereas<br />

the block<strong>in</strong>g <strong>of</strong> migration by sGC <strong>in</strong>hibitor was reversed <strong>in</strong> the presence <strong>of</strong> a cell membrane<br />

permeable analogue <strong>of</strong> cGMP. These results strongly <strong>in</strong>dicate that the NO/cGMP/PKG signal<strong>in</strong>g<br />

pathway is essential for the migration <strong>of</strong> human neuronal precursor cells. Moreover, exogenous use<br />

<strong>of</strong> NO and membrane permeable analogue <strong>of</strong> cGMP facilitated cell motility. Although sGC and<br />

PKG was identified to act downstream <strong>of</strong> NO, I did not exam<strong>in</strong>e how the NO activated cGMP and<br />

PKG modulate neuronal migration. S<strong>in</strong>ce one common feature among signal<strong>in</strong>g pathways that<br />

regulate neuronal migration is the eventual <strong>in</strong>volvement <strong>of</strong> the cytoskeleton, I predict that NO<br />

activated cGMP/PKG pathway may cause reorganization <strong>of</strong> act<strong>in</strong>. The reorganization <strong>of</strong> act<strong>in</strong><br />

cytoskeleton <strong>in</strong> response to cGMP/PKG that <strong>in</strong>volves RhoA GTPase and phosphorylation <strong>of</strong><br />

Enabled/vasodilator-stimulated phosphoprote<strong>in</strong> family prote<strong>in</strong>s has been reported (Sporbert et al.,<br />

1999; Gudi et al., 2002; Borán and García 2007; L<strong>in</strong>dsay et al., 2007, Zulauf et al., 2009).<br />

S<strong>in</strong>ce several studies implicated that NO <strong>in</strong>hibit neuronal precursor cells proliferation while<br />

enhanc<strong>in</strong>g neuronal differentiation (Ghigo et al., 1998; Phung et al., 1999; Cheng, 2003; Ciani et<br />

12


al., 2004; 2006), I tested the effect <strong>of</strong> chemicals on neuronal proliferation and differentiation. Here,<br />

concentration <strong>of</strong> both activators and <strong>in</strong>hibitors <strong>of</strong> small bioactive chemicals used to modulate cell<br />

migration did not significantly affect cell viability, proliferation and neuronal differentiation<br />

<strong>in</strong>dicat<strong>in</strong>g that the observed effect was ma<strong>in</strong>ly on cell migration. However, the NO donor (100 µM<br />

NOC-18) differentially coord<strong>in</strong>ated cell motility and proliferation. At this concentration, NOC-18<br />

seemed to facilitate cell migration while <strong>in</strong>hibit<strong>in</strong>g proliferation (Fig. 6, Tegenge and Bicker, 2009).<br />

This effect <strong>of</strong> NOC-18 on neuronal progenitor proliferation is <strong>in</strong> agreement with several studies that<br />

showed anti-proliferative effect <strong>of</strong> NO (Enikolopov et al., 1999; Murillo-Carretero et al, 2002;<br />

Ciani et al., 2004;Peunova et al., 2007). Moreover, we observed that under <strong>in</strong>hibition <strong>of</strong> nNOS<br />

there was a modest decrease <strong>in</strong> the percentage <strong>of</strong> neuronal cells while the cGMP analogue slightly<br />

<strong>in</strong>creased differentiation <strong>in</strong>to neurons (Tegenge and Bicker, 2009). Thus, the present data show<br />

concentration dependent effects <strong>of</strong> NO/cGMP signal<strong>in</strong>g on the dist<strong>in</strong>ct process <strong>of</strong> neuronal<br />

proliferation, migration and differentiation which could account partly for the reported<br />

<strong>in</strong>consistency on the role <strong>of</strong> NO at early stage <strong>of</strong> nervous system development.<br />

In summary, this study presented a model <strong>of</strong> human neuronal precursor cells as three-dimensional<br />

spherical aggregates that proliferate, migrate and differentiate <strong>in</strong>to neurons. The migration <strong>of</strong> cells<br />

from NT2 spherical aggregates is modulated by NO/cGMP/PKG signal<strong>in</strong>g pathway. S<strong>in</strong>ce NT2<br />

cells spherical aggregates elaborate neuronal process, it will be <strong>in</strong>terest<strong>in</strong>g to <strong>in</strong>vestigate the effect<br />

<strong>of</strong> NO on the growth cone response and neurite outgrowth.<br />

2.2. <strong>Nitric</strong> oxide and cyclic nucleotide signal transduction modulates synaptic vesicle turnover<br />

<strong>in</strong> human model neurons<br />

The human embryonal carc<strong>in</strong>oma stem cell l<strong>in</strong>e Ntera-2/cl.D1 (NT2), is extensively studied as a<br />

model <strong>of</strong> human neuronal differentiation <strong>in</strong> vitro. The NT2 cells have been demonstrated to<br />

term<strong>in</strong>ally differentiate <strong>in</strong>to functional postmitotic neurons <strong>in</strong> vitro (Andrews, 1984; Pleasure et al.,<br />

1992; Paquet-Durand and Bicker 2007). The major drawback was the rather lengthy differentiation<br />

protocol, which has been shortened significantly <strong>in</strong> recent years by employ<strong>in</strong>g the aggregate culture<br />

method (Paquet-Durand et al., 2003; Podrygajlo et al., 2009). In this study we followed the<br />

synaptic maturation <strong>of</strong> human NT2 neurons generated by this novel culture method. <strong>Human</strong> NT2<br />

neurons express the dendritic marker (MAP2) and a marker <strong>of</strong> axogenesis (phosphorylated Tau).<br />

While MAP2 is expressed <strong>in</strong> short process at all stages <strong>of</strong> NT2 neurons, the sta<strong>in</strong><strong>in</strong>g for Tau on<br />

longer neurites appeared only later upon cultur<strong>in</strong>g NT2 neurons (about 4 weeks).<br />

Dur<strong>in</strong>g synaptogenesis the amount <strong>of</strong> both pre- and postsynaptic prote<strong>in</strong>s <strong>in</strong>creases considerably<br />

(McAllister A.K, 2007). Here, synaptic maturation was followed by immunocytochemical sta<strong>in</strong><strong>in</strong>g<br />

13


<strong>of</strong> the pre-synaptic vesicle associated prote<strong>in</strong>s (synaps<strong>in</strong> I and synaptotagm<strong>in</strong> I). The sta<strong>in</strong><strong>in</strong>g <strong>of</strong><br />

both prote<strong>in</strong>s appeared as punctate along the neurites and <strong>in</strong>creased significantly with<strong>in</strong> 2-4 weeks<br />

<strong>of</strong> <strong>in</strong> vitro culture <strong>in</strong>dicative <strong>of</strong> pre-synaptic maturation (Figs 1&2, Tegenge et al., 2009). Evidence<br />

for activity-dependent pre-synaptic vesicle release was provided by employ<strong>in</strong>g antibody aga<strong>in</strong>st the<br />

lum<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> synaptotagm<strong>in</strong> I. Upon fusion <strong>of</strong> synaptic vesicles with the plasma membrane,<br />

the lum<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> synaptotagm<strong>in</strong> becomes exposed to the neuronal surface, allow<strong>in</strong>g the<br />

b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> the antibody and their subsequent <strong>in</strong>ternalization through the endocytosis <strong>of</strong> synaptic<br />

vesicle (Figure 2a). In human NT2 neurons I demonstrated that lum<strong>in</strong>al synaptotagm<strong>in</strong> antibody is<br />

<strong>in</strong>corporated <strong>in</strong> both depolarization and calcium dependent manner, which suggest that the neurons<br />

display synaptic vesicle recycl<strong>in</strong>g. In NT2 neurons cultured only for 1 week we hardly detected<br />

lum<strong>in</strong>al synaptotagm<strong>in</strong> punctate sta<strong>in</strong><strong>in</strong>g compared to <strong>in</strong>tense sta<strong>in</strong><strong>in</strong>g with<strong>in</strong> 2-4 weeks. Thus,<br />

activity-dependent <strong>in</strong>corporation <strong>of</strong> lum<strong>in</strong>al synaptotagm<strong>in</strong> antibody depends on the length <strong>of</strong> <strong>in</strong><br />

vitro culture <strong>of</strong> the neurons <strong>in</strong>dicat<strong>in</strong>g that human NT2 neurons undergo synaptic maturation<br />

(Tegenge et al., 2009).<br />

Figure 2. Methods to monitor the synaptic vesicle recycl<strong>in</strong>g and exocytosis <strong>in</strong> develop<strong>in</strong>g neurons.<br />

(a) Use <strong>of</strong> antibodies aga<strong>in</strong>st the lum<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> the synaptic vesicle prote<strong>in</strong> synaptotagm<strong>in</strong> (Syt-<br />

14


ecto abs). Upon fusion <strong>of</strong> synaptic vesicles with the plasma membrane, the lum<strong>in</strong>al doma<strong>in</strong> <strong>of</strong><br />

synaptotagm<strong>in</strong> becomes exposed to the neuronal surface, allow<strong>in</strong>g the b<strong>in</strong>d<strong>in</strong>g <strong>of</strong> Syt-ecto abs and<br />

their subsequent <strong>in</strong>ternalization through synaptic-vesicle endocytosis. (b) Use <strong>of</strong> the amphiphilic<br />

dye FM1–43, which very effectively labels endocytic vesicles (load<strong>in</strong>g and wash) and dissociates<br />

from membranes upon re-exposure <strong>of</strong> the vesicle lumen to the extracellular medium dur<strong>in</strong>g<br />

exocytosis (unload<strong>in</strong>g). Adopted from (Matteoli et al., 2004).<br />

Independently, a second functional evidence for human NT2 neurons was provided by imag<strong>in</strong>g<br />

synaptic vesicle recycl<strong>in</strong>g and exocytosis employ<strong>in</strong>g fluorescent stryl dye, FM1-43. The dye labels<br />

synaptic vesicle that undergo vesicle recycl<strong>in</strong>g and dissociates from the vesicles to the extracellular<br />

medium upon a second round <strong>of</strong> exocytosis (Figure 2b). FM1-43 imag<strong>in</strong>g revealed that depolarized<br />

NT2 neurons display synaptic vesicle recycl<strong>in</strong>g and exocytosis <strong>in</strong> the presence <strong>of</strong> calcium. This pre-<br />

synaptic vesicle exocytosis <strong>in</strong>duced by high K + presumably <strong>in</strong>dicates neurotransmitter release at the<br />

active zone <strong>of</strong> nerve term<strong>in</strong>als. Recently, our group has shown that <strong>in</strong> the absence <strong>of</strong> glial cells, NT2<br />

neurons showed spontaneous synaptic currents and responded to exogenous application <strong>of</strong> GABA<br />

and glutamate neurotransmitters (Podrygajlo et al., 2009, Podrygajlo et al., 2010). These results<br />

suggest that GABA and glutamate are putatively released upon high K + stimulation. Thus, NT2<br />

neurons represent pure human neuronal culture that displays rapid vesicle exocytosis. As a result,<br />

NT2 neurons can be utilized as human model neurons to understand the cellular and molecular basis<br />

<strong>of</strong> pre-synaptic vesicle release. Furthermore, time dependent synaptic maturation <strong>of</strong> NT2 neurons<br />

can be exploited to screen chemical compounds <strong>in</strong>fluenc<strong>in</strong>g the expression <strong>of</strong> pre-synaptic prote<strong>in</strong>s<br />

and pre-synaptic release. In this regard NT2 neurons have greater potential as compared to the<br />

widely used the neuroendocr<strong>in</strong>e cell l<strong>in</strong>e PC12 cells (Wester<strong>in</strong>k and Ew<strong>in</strong>g, 2008) not only because<br />

<strong>of</strong> its human orig<strong>in</strong> but also due to its term<strong>in</strong>al differentiation <strong>in</strong>to functional neurons.<br />

Numerous studies demonstrated that NO <strong>in</strong>duces pre-synaptic vesicle release (Hawk<strong>in</strong>s et al., 1994;<br />

Arancio et al., 1996; Meffert et al., 1996; Sporns and Jenk<strong>in</strong>son, 1997; Wildemann and Bicker,<br />

1999; Li et al., 2002). Furthermore, NO/cGMP pathway has been shown to regulate synaptic vesicle<br />

endocytosis and recycl<strong>in</strong>g by <strong>in</strong>creas<strong>in</strong>g pre-synaptic phosphatidyl<strong>in</strong>ositol 4, 5-bisphosphate (PIP2)<br />

(Micheva et al. 2003; Petrov et al., 2008). In this study we showed for the first time that NO/cGMP<br />

signal<strong>in</strong>g cascade <strong>in</strong>duces synaptic vesicle recycl<strong>in</strong>g and pre-synaptic release <strong>in</strong> human model<br />

neurons. Firstly, expression <strong>of</strong> nNOS and NO-sensitive sGC was demonstrated suggest<strong>in</strong>g<br />

functional NO/cGMP signal<strong>in</strong>g is present <strong>in</strong> human NT2 neurons. Previously, it has been shown<br />

that expression <strong>of</strong> NO-sensitive sGC correlates with synaptogenesis (Truman et al, 1996; Ball and<br />

Truman, 1998; Bicker, 2005). Secondly, two NO donors and membrane permeable cGMP analogue<br />

facilitate synaptic vesicle recycl<strong>in</strong>g and vesicle exocytosis <strong>in</strong> human NT2 neurons (Tegenge et al.,<br />

15


2009). In develop<strong>in</strong>g neurons synaptic vesicle turnover is important <strong>in</strong> the assembly <strong>of</strong> new synaptic<br />

sites (Matteoli et al., 2004). Moreover, activation <strong>of</strong> NO signal<strong>in</strong>g has been demonstrated to<br />

facilitate activity-dependent pre-synaptic remodell<strong>in</strong>g that contributes to synaptogenesis<br />

(Nikonenko et al., 2003; 2008). Thus, <strong>in</strong>duction <strong>of</strong> synaptic turnover by NO donors and cGMP<br />

analogue <strong>in</strong> human NT2 neurons suggest that NO signal<strong>in</strong>g may <strong>in</strong>volve <strong>in</strong> activity-dependent<br />

synaptic maturation <strong>of</strong> human nerve cells.<br />

The data further demonstrated that the cAMP/PKA signal<strong>in</strong>g <strong>in</strong>duces synaptic vesicle turnover <strong>in</strong><br />

human NT2 neurons, which possibly <strong>in</strong>volve phosporylation <strong>of</strong> synaps<strong>in</strong>. Phosphorylation <strong>of</strong><br />

synaps<strong>in</strong> by the cAMP/PKA signal<strong>in</strong>g that leads to the release <strong>of</strong> vesicles from the reserve pool has<br />

been suggested to enhance pre-synaptic vesicle releases (Fiumara et al., 2004; 2007; Bonanomi et<br />

al., 2005; Menegon et al., 2006).<br />

Taken together, I have shown that human NT2 neurons undergo pre-synaptic maturation <strong>in</strong> vitro.<br />

Moreover, the release <strong>of</strong> neurotransmitter at pre-synaptic term<strong>in</strong>als <strong>of</strong> human NT2 neurons could be<br />

modulated by NO and cyclic nucleotide signal<strong>in</strong>g cascade. These result po<strong>in</strong>ts toward critical role<br />

<strong>of</strong> NO/cyclic nucleotide signal<strong>in</strong>g as a regulator <strong>of</strong> neurotransmitter release <strong>in</strong> develop<strong>in</strong>g human<br />

nerve cells.<br />

2.3. <strong>Nitric</strong> oxide signal<strong>in</strong>g as regulator <strong>of</strong> human neuronal progenitor cell migration<br />

Previously, we demonstrated that NO/cGMP signal<strong>in</strong>g facilitates the migration <strong>of</strong> cultured human<br />

neuronal precursor (NT2) cells (Tegenge and Bicker, 2009). S<strong>in</strong>ce the NT2 cell l<strong>in</strong>e was obta<strong>in</strong>ed<br />

from a human teratocarc<strong>in</strong>oma (Andrews, 1984) and we cannot rule out effects <strong>of</strong> genetic<br />

rearrangements <strong>in</strong> the cancer cells, it is problematic to relate these f<strong>in</strong>d<strong>in</strong>gs to normal human neural<br />

development. Multi-potent human neuronal progenitor cells can be obta<strong>in</strong>ed from cells <strong>of</strong><br />

ectodermal l<strong>in</strong>eage and are restricted <strong>in</strong> fate to develop <strong>in</strong>to neurons, astrocytes and<br />

oligodendrocytes. Stem cells can also be isolated from both develop<strong>in</strong>g human bra<strong>in</strong> and restricted<br />

areas <strong>of</strong> adult human bra<strong>in</strong>. Studies over the past decade have focused ma<strong>in</strong>ly on the use <strong>of</strong> stem<br />

cells <strong>in</strong> transplantation therapy for various diseases models. However, stem cells derived neuronal<br />

progenitor cells have the potential to <strong>in</strong>vestigate the molecular and cellular basis <strong>of</strong> human bra<strong>in</strong><br />

development.<br />

In this study, we used fetal human neural progenitor cells (hNPCs) cultured as neurospheres to<br />

<strong>in</strong>vestigate the role <strong>of</strong> NO/cGMP signal<strong>in</strong>g <strong>in</strong> neuronal migration. hNPCs derive from 16 to 20week<br />

old female or male fetal whole bra<strong>in</strong> homogenates, which have the capacity to differentiate<br />

<strong>in</strong>to neuronal and glial cells (Moors et al., 2009; Breier, 2009). Cultur<strong>in</strong>g proliferat<strong>in</strong>g human<br />

neurospheres on lam<strong>in</strong><strong>in</strong> coated adhesive substrates resulted <strong>in</strong> cells that migrate out <strong>of</strong> the spheres<br />

16


which were accompanied by differentiation <strong>in</strong>to both neuronal and glial cells. Intrigu<strong>in</strong>gly, cells that<br />

migrate out respond to exogenous stimulation <strong>of</strong> NO and synthesize <strong>in</strong>creas<strong>in</strong>g level <strong>of</strong> cGMP.<br />

These results provide first experimental evidence that functional NO-sensitive sGC is expressed<br />

early dur<strong>in</strong>g human bra<strong>in</strong> development. The cGMP positive cells at the forefront <strong>of</strong> migration were<br />

co-sta<strong>in</strong>ed for either nest<strong>in</strong> or GFAP but not for ����-tubul<strong>in</strong>. Thus, compared to NT2 cell culture<br />

that lack GFAP positive cells, the hNPCs may mimic aspects <strong>of</strong> radial neuroblast migration dur<strong>in</strong>g<br />

human bra<strong>in</strong> development.<br />

Application <strong>of</strong> small bioactive enzyme <strong>in</strong>hibitors and activators <strong>in</strong> human neurosphere culture<br />

implicate that NO/cGMP/PKG signal<strong>in</strong>g is key regulator <strong>of</strong> neuronal progenitor migration.<br />

Compared to the NT2 cell culture, <strong>in</strong> hNPCs <strong>in</strong>verted U-shaped dose-response curve for the NO<br />

donor NOC-18 was obta<strong>in</strong>ed at relatively smaller dose range (1-100 µM). Our data demonstrate that<br />

at 10 µM NOC-18 appeared to facilitate neuronal progenitor cell migration while at relatively<br />

higher concentration (100 µM) <strong>in</strong>hibition <strong>of</strong> cell migration was observed. These result suggest that<br />

NO could play dual role on neuronal migration depend<strong>in</strong>g on the concentration. Such dual role <strong>of</strong><br />

NO on cell migration and neurite outgrowth that depends on concentration has been reported<br />

previously (Elfer<strong>in</strong>k, and VanUffelen, 1996; Trimm and Rehder, 2004). Furthermore, we directly<br />

showed the <strong>in</strong>volvement <strong>of</strong> sGC and PKG <strong>in</strong> NO <strong>in</strong>duced facilitation <strong>of</strong> hNPCs migration. Here, coapplication<br />

<strong>of</strong> NO donor with enzyme <strong>in</strong>hibitors <strong>of</strong> sGC or PKG resulted <strong>in</strong> blocked <strong>of</strong> NO <strong>in</strong>duced<br />

facilitation <strong>of</strong> hNPCs migration suggest<strong>in</strong>g that low concentration <strong>of</strong> NO facilitates neuronal<br />

progenitor migration via sGC and PKG signal<strong>in</strong>g pathway.<br />

In summary, this study provides the first evidence for the <strong>in</strong>volvement <strong>of</strong> NO signal<strong>in</strong>g <strong>in</strong> human<br />

neuronal development. S<strong>in</strong>ce neuronal migration is controlled by a complex comb<strong>in</strong>ation <strong>of</strong><br />

chemical signals, it would be <strong>in</strong>terest<strong>in</strong>g to comprehend how the NO signal regulates neuronal<br />

migration <strong>in</strong> comb<strong>in</strong>ation with other known signals. Understand<strong>in</strong>g the cellular and molecular<br />

mechanisms <strong>of</strong> neuronal migration may <strong>of</strong>fer hopes to develop new therapeutic agents for human<br />

congential disorders associated with defective neuronal migration. Our f<strong>in</strong>d<strong>in</strong>gs have further<br />

implications <strong>in</strong> adult neurogenesis. In recent years research <strong>in</strong> NSC biology focus on restoration <strong>of</strong><br />

damaged neural networks <strong>in</strong> the various neurodegenerative diseases by promot<strong>in</strong>g endogenous<br />

neurogenesis or transplantation <strong>of</strong> NSC. To this end, manipulation <strong>of</strong> NO/cGMP signal<strong>in</strong>g may<br />

facilitate migration and <strong>in</strong>tegration <strong>of</strong> endogenous or transplanted neuronal progenitor cells <strong>in</strong>to the<br />

exist<strong>in</strong>g neuronal network which could promote functional recovery <strong>in</strong> various neurodegenerative<br />

diseases.<br />

17


References<br />

Agulló, L., and García, A. (1992). Different receptors mediate stimulation <strong>of</strong> nitric oxide-dependent<br />

cyclic GMP formation <strong>in</strong> neurons and astrocytes <strong>in</strong> culture. Biochem. Biophys. Res.<br />

Commun 182, 1362-1368.<br />

Andrews, P. W. (1984). Ret<strong>in</strong>oic acid <strong>in</strong>duces neuronal differentiation <strong>of</strong> a cloned human<br />

embryonal carc<strong>in</strong>oma cell l<strong>in</strong>e <strong>in</strong> vitro. Dev. Biol 103, 285-293.<br />

Arnold, W. P., Mittal, C. K., Katsuki, S., and Murad, F. (1977). <strong>Nitric</strong> oxide activates guanylate<br />

cyclase and <strong>in</strong>creases guanos<strong>in</strong>e 3':5'-cyclic monophosphate levels <strong>in</strong> various tissue<br />

preparations. Proc. Natl. Acad. Sci. U.S.A 74, 3203-3207.<br />

Ayala, R., Shu, T., and Tsai, L. (2007). Trekk<strong>in</strong>g across the bra<strong>in</strong>: the journey <strong>of</strong> neuronal<br />

migration. Cell 128, 29-43.<br />

Ball, E. E., and Truman, J. W. (1998). Develop<strong>in</strong>g grasshopper neurons show variable levels <strong>of</strong><br />

guanylyl cyclase activity on arrival at their targets. J. Comp. Neurol 394, 1-13.<br />

Baltrons, M. A., Pedraza, C., Sardón, T., Navarra, M., and García, A. (2003). Regulation <strong>of</strong> NOdependent<br />

cyclic GMP formation by <strong>in</strong>flammatory agents <strong>in</strong> neural cells. Toxicol. Lett 139,<br />

191-198.<br />

Bicker, G. (2001). <strong>Nitric</strong> oxide: an unconventional messenger <strong>in</strong> the nervous system <strong>of</strong> an<br />

orthopteroid <strong>in</strong>sect. Arch. Insect Biochem. Physiol 48, 100-110.<br />

Bicker, G. (2005). STOP and GO with NO: nitric oxide as a regulator <strong>of</strong> cell motility <strong>in</strong> simple<br />

bra<strong>in</strong>s. Bioessays 27, 495-505.<br />

Blaise, G. A., Gauv<strong>in</strong>, D., Gangal, M., and Authier, S. (2005). <strong>Nitric</strong> oxide, cell signal<strong>in</strong>g and cell<br />

death. Toxicology 208, 177-192.<br />

Boehn<strong>in</strong>g, D., and Snyder, S. H. (2003). Novel neural modulators. Annu. Rev. Neurosci 26, 105-<br />

131.<br />

Bonanomi, D., Menegon, A., Miccio, A., Ferrari, G., Corradi, A., Kao, H., Benfenati, F., and<br />

Valtorta, F. (2005). Phosphorylation <strong>of</strong> synaps<strong>in</strong> I by cAMP-dependent prote<strong>in</strong> k<strong>in</strong>ase<br />

controls synaptic vesicle dynamics <strong>in</strong> develop<strong>in</strong>g neurons. J. Neurosci 25, 7299-7308.<br />

Borán, M. S., and García, A. (2007). The cyclic GMP-prote<strong>in</strong> k<strong>in</strong>ase G pathway regulates<br />

cytoskeleton dynamics and motility <strong>in</strong> astrocytes. J. Neurochem 102, 216-230.<br />

Boucherie, C., and Hermans, E. (2009). Adult stem cell therapies for neurological disorders:<br />

benefits beyond neuronal replacement? J. Neurosci. Res 87, 1509-1521.<br />

Bredt, D. S., Hwang, P. M., and Snyder, S. H. (1990). Localization <strong>of</strong> nitric oxide synthase<br />

<strong>in</strong>dicat<strong>in</strong>g a neural role for nitric oxide. Nature 347, 768-770.<br />

Bredt, D. S., and Snyder, S. H. (1989). <strong>Nitric</strong> oxide mediates glutamate-l<strong>in</strong>ked enhancement <strong>of</strong><br />

cGMP levels <strong>in</strong> the cerebellum. Proc. Natl. Acad. Sci. U.S.A 86, 9030-9033.<br />

Bredt, D. S., and Snyder, S. H. (1994). Transient nitric oxide synthase neurons <strong>in</strong> embryonic<br />

cerebral cortical plate, sensory ganglia, and olfactory epithelium. Neuron 13, 301-313.<br />

18


Breier, J. M., Gassmann, K., Kayser, R., Stegeman, H., De Groot, D., Fritsche, E., and Shafer, T. J.<br />

(2009). Neural progenitor cells as models for high-throughput screens <strong>of</strong> developmental<br />

neurotoxicity: State <strong>of</strong> the science. Neurotoxicol Teratol (In press).<br />

Bulseco, D. A., Poluha, W., Schonh<strong>of</strong>f, C. M., Daou, M. C., Condon, P. J., and Ross, A. H. (2001).<br />

Cell-cycle arrest <strong>in</strong> TrkA-express<strong>in</strong>g NIH3T3 cells <strong>in</strong>volves nitric oxide synthase. J. Cell.<br />

Biochem 81, 193-204.<br />

Cárdenas, A., Moro, M. A., Hurtado, O., Leza, J. C., and Lizasoa<strong>in</strong>, I. (2005). Dual role <strong>of</strong> nitric<br />

oxide <strong>in</strong> adult neurogenesis. Bra<strong>in</strong> Res. Bra<strong>in</strong> Res. Rev 50, 1-6.<br />

Castro-Blanco, S., Enc<strong>in</strong>as, J. M., Serrano, J., Alonso, D., Gómez, M. B., Sánchez, J., Ríos-Tejada,<br />

F., Fernández-Vizarra, P., Fernández, A. P., Martínez-Murillo, R., et al. (2003). Expression<br />

<strong>of</strong> nitrergic system and prote<strong>in</strong> nitration <strong>in</strong> adult rat bra<strong>in</strong>s submitted to acute hypobaric<br />

hypoxia. <strong>Nitric</strong> <strong>Oxide</strong> 8, 182-201.<br />

Chen, J., Tu, Y., Moon, C., Matarazzo, V., Palmer, A. M., and Ronnett, G. V. (2004). The<br />

localization <strong>of</strong> neuronal nitric oxide synthase may <strong>in</strong>fluence its role <strong>in</strong> neuronal precursor<br />

proliferation and synaptic ma<strong>in</strong>tenance. Dev. Biol 269, 165-182.<br />

Cheng, A., Wang, S., Cai, J., Rao, M. S., and Mattson, M. P. (2003). <strong>Nitric</strong> oxide acts <strong>in</strong> a positive<br />

feedback loop with BDNF to regulate neural progenitor cell proliferation and differentiation<br />

<strong>in</strong> the mammalian bra<strong>in</strong>. Dev. Biol 258, 319-333.<br />

Cho, D., Nakamura, T., Fang, J., Cieplak, P., Godzik, A., Gu, Z., and Lipton, S. A. (2009). Snitrosylation<br />

<strong>of</strong> Drp1 mediates beta-amyloid-related mitochondrial fission and neuronal<br />

<strong>in</strong>jury. Science 324, 102-105.<br />

Ciani, E., Calvanese, V., Crochemore, C., Bartesaghi, R., and Contestabile, A. (2006). Proliferation<br />

<strong>of</strong> cerebellar precursor cells is negatively regulated by nitric oxide <strong>in</strong> newborn rat. J. Cell.<br />

Sci 119, 3161-3170.<br />

Ciani, E., Severi, S., Contestabile, A., Bartesaghi, R., and Contestabile, A. (2004). <strong>Nitric</strong> oxide<br />

negatively regulates proliferation and promotes neuronal differentiation through N-Myc<br />

downregulation. J. Cell. Sci 117, 4727-4737.<br />

Contestabile, A. (2008). Regulation <strong>of</strong> transcription factors by nitric oxide <strong>in</strong> neurons and <strong>in</strong> neuralderived<br />

tumor cells. Prog. Neurobiol 84, 317-328.<br />

Corb<strong>in</strong>, J. G., Gaiano, N., Juliano, S. L., Poluch, S., Stancik, E., and Haydar, T. F. (2008).<br />

Regulation <strong>of</strong> neural progenitor cell development <strong>in</strong> the nervous system. J. Neurochem 106,<br />

2272-2287.<br />

Crosbie, R. H., Straub, V., Yun, H. Y., Lee, J. C., Rafael, J. A., Chamberla<strong>in</strong>, J. S., Dawson, V. L.,<br />

Dawson, T. M., and Campbell, K. P. (1998). mdx muscle pathology is <strong>in</strong>dependent <strong>of</strong> nNOS<br />

perturbation. Hum. Mol. Genet 7, 823-829.<br />

Cui, X., Chen, J., Zacharek, A., Roberts, C., Yang, Y., and Chopp, M. (2009). <strong>Nitric</strong> oxide donor<br />

up-regulation <strong>of</strong> SDF1/CXCR4 and Ang1/Tie2 promotes neuroblast cell migration after<br />

stroke. J. Neurosci. Res 87, 86-95.<br />

Curtis, M. A., Connor, B., and Faull, R. L. M. (2003). Neurogenesis <strong>in</strong> the diseased adult human<br />

19


a<strong>in</strong>--new therapeutic strategies for neurodegenerative diseases. Cell Cycle 2, 428-430.<br />

De Palma, C., Falcone, S., Panzeri, C., Radice, S., Bassi, M. T., and Clementi, E. (2008).<br />

Endothelial nitric oxide synthase overexpression by neuronal cells <strong>in</strong> neurodegeneration: a<br />

l<strong>in</strong>k between <strong>in</strong>flammation and neuroprotection. J. Neurochem 106, 193-204.<br />

Di Matteo, V., Pierucci, M., Benigno, A., Crescimanno, G., Esposito, E., and Di Giovanni, G.<br />

(2009). Involvement <strong>of</strong> nitric oxide <strong>in</strong> nigrostriatal dopam<strong>in</strong>ergic system degeneration: a<br />

neurochemical study . Ann. N. Y. Acad. Sci 1155, 309-315.<br />

Donovan, P. J., and Gearhart, J. (2001). The end <strong>of</strong> the beg<strong>in</strong>n<strong>in</strong>g for pluripotent stem cells. Nature<br />

414, 92-97.<br />

Enikolopov, G., Banerji, J., and Kuz<strong>in</strong>, B. (1999). <strong>Nitric</strong> oxide and Drosophila development. Cell<br />

Death Differ 6, 956-963.<br />

Elfer<strong>in</strong>k, J.G., and VanUffelen, B.E. (1996). The role <strong>of</strong> cyclic nucleotides <strong>in</strong> neutrophil migration.<br />

Gen Pharmaco 27: 387-393.<br />

Ernst, A. F., Gallo, G., Letourneau, P. C., and McLoon, S. C. (2000). Stabilization <strong>of</strong> grow<strong>in</strong>g<br />

ret<strong>in</strong>al axons by the comb<strong>in</strong>ed signal<strong>in</strong>g <strong>of</strong> nitric oxide and bra<strong>in</strong>-derived neurotrophic<br />

factor. J. Neurosci 20, 1458-1469.<br />

Estrada, C., and Murillo-Carretero, M. (2005). <strong>Nitric</strong> oxide and adult neurogenesis <strong>in</strong> health and<br />

disease. Neuroscientist 11, 294-307.<br />

Evangelopoulos, M.E., Wüller, S., Weis, J., and Krüttgen A.(2010). A role <strong>of</strong> nitric oxide <strong>in</strong> neurite<br />

outgrowth <strong>of</strong> neuroblastoma cells triggered by mevastat<strong>in</strong> or serum reduction. Neurosci. Lett<br />

468, 28-33.<br />

Fischer, A., Sananbenesi, F., Wang, X., Dobb<strong>in</strong>, M., and Tsai, L. (2007). Recovery <strong>of</strong> learn<strong>in</strong>g and<br />

memory is associated with chromat<strong>in</strong> remodell<strong>in</strong>g. Nature 447, 178-182.<br />

Fiumara, F., Giovedì, S., Menegon, A., Milanese, C., Merlo, D., Montarolo, P. G., Valtorta, F.,<br />

Benfenati, F., and Ghirardi, M. (2004). Phosphorylation by cAMP-dependent prote<strong>in</strong> k<strong>in</strong>ase<br />

is essential for synaps<strong>in</strong>-<strong>in</strong>duced enhancement <strong>of</strong> neurotransmitter release <strong>in</strong> <strong>in</strong>vertebrate<br />

neurons. J. Cell. Sci 117, 5145-5154.<br />

Fiumara, F., Milanese, C., Corradi, A., Giovedì, S., Leit<strong>in</strong>ger, G., Menegon, A., Montarolo, P. G.,<br />

Benfenati, F., and Ghirardi, M. (2007). Phosphorylation <strong>of</strong> synaps<strong>in</strong> doma<strong>in</strong> A is required<br />

for post-tetanic potentiation. J. Cell. Sci 120, 3228-3237.<br />

Förstermann, U., Gath, I., Schwarz, P., Closs, E. I., and Kle<strong>in</strong>ert, H. (1995). Is<strong>of</strong>orms <strong>of</strong> nitric oxide<br />

synthase. Properties, cellular distribution and expressional control. Biochem. Pharmacol 50,<br />

1321-1332.<br />

Foster, J. A., and Phelps, P. E. (2000). Neurons express<strong>in</strong>g NADPH-diaphorase <strong>in</strong> the develop<strong>in</strong>g<br />

human sp<strong>in</strong>al cord. J. Comp. Neurol 427, 417-427.<br />

Friebe, A, and Koesl<strong>in</strong>g, D. (2009). The function <strong>of</strong> NO-sensitive guanylyl cyclase: what we can<br />

learn from genetic mouse models. <strong>Nitric</strong> <strong>Oxide</strong> 21, 149-156.<br />

Fritsche, E., Cl<strong>in</strong>e, J. E., Nguyen, N., Scanlan, T. S., and Abel, J. (2005). Polychlor<strong>in</strong>ated biphenyls<br />

20


disturb differentiation <strong>of</strong> normal human neural progenitor cells: clue for <strong>in</strong>volvement <strong>of</strong><br />

thyroid hormone receptors. Environ. Health Perspect 113, 871-876.<br />

Fritzen, S., Schmitt, A., Köth, K., Sommer, C., Lesch, K., and Reif, A. (2007). Neuronal nitric<br />

oxide synthase (NOS-I) knockout <strong>in</strong>creases the survival rate <strong>of</strong> neural cells <strong>in</strong> the<br />

hippocampus <strong>in</strong>dependently <strong>of</strong> BDNF. Mol. Cell. Neurosci 35, 261-271.<br />

Furchgott, R. F., and Zawadzki, J. V. (1980). The obligatory role <strong>of</strong> endothelial cells <strong>in</strong> the<br />

relaxation <strong>of</strong> arterial smooth muscle by acetylchol<strong>in</strong>e. Nature 288, 373-376.<br />

Gage, F. H. (2000). Mammalian neural stem cells. Science 287, 1433-1438.<br />

Garthwaite, J., Charles, S. L., and Chess-Williams, R. (1988). Endothelium-derived relax<strong>in</strong>g factor<br />

release on activation <strong>of</strong> NMDA receptors suggests role as <strong>in</strong>tercellular messenger <strong>in</strong> the<br />

bra<strong>in</strong>. Nature 336, 385-388.<br />

Garthwaite, J. (2008). Concepts <strong>of</strong> neural nitric oxide-mediated transmission. Eur. J. Neurosci 27,<br />

2783-2802.<br />

Ghigo, D., Priotto, C., Miglior<strong>in</strong>o, D., Gerom<strong>in</strong>, D., Franch<strong>in</strong>o, C., Todde, R., Costamagna, C.,<br />

Pescarmona, G., and Bosia, A. (1998). Ret<strong>in</strong>oic acid-<strong>in</strong>duced differentiation <strong>in</strong> a human<br />

neuroblastoma cell l<strong>in</strong>e is associated with an <strong>in</strong>crease <strong>in</strong> nitric oxide synthesis. J. Cell.<br />

Physiol 174, 99-106.<br />

Gibbs, S. M., and Truman, J. W. (1998). <strong>Nitric</strong> oxide and cyclic GMP regulate ret<strong>in</strong>al pattern<strong>in</strong>g <strong>in</strong><br />

the optic lobe <strong>of</strong> Drosophila. Neuron 20, 83-93.<br />

Gibbs, S. M. (2003). Regulation <strong>of</strong> neuronal proliferation and differentiation by nitric oxide. Mol.<br />

Neurobiol 27, 107-120.<br />

Goldberg, D. J., and Burmeister, D. W. (1989). Look<strong>in</strong>g <strong>in</strong>to growth cones. Trends Neurosci 12,<br />

503-506.<br />

Grzybicki, D., Gebhart, G. F., and Murphy, S. (1996). Expression <strong>of</strong> nitric oxide synthase type II <strong>in</strong><br />

the sp<strong>in</strong>al cord under conditions produc<strong>in</strong>g thermal hyperalgesia. J. Chem. Neuroanat 10,<br />

221-229.<br />

Gudi, T., Chen, J. C., Casteel, D. E., Seasholtz, T. M., Boss, G. R., and Pilz, R. B. (2002). cGMPdependent<br />

prote<strong>in</strong> k<strong>in</strong>ase <strong>in</strong>hibits serum-response element-dependent transcription by<br />

<strong>in</strong>hibit<strong>in</strong>g rho activation and functions. J. Biol. Chem 277, 37382-37393.<br />

Guillema<strong>in</strong>, I., Alonso, G., Patey, G., Privat, A., and Chaudieu, I. (2000). <strong>Human</strong> NT2 neurons<br />

express a large variety <strong>of</strong> neurotransmission phenotypes <strong>in</strong> vitro. J. Comp. Neurol 422, 380-<br />

395.<br />

Gutièrrez-Mec<strong>in</strong>as, M., Crespo, C., Blasco-Ibáñez, J. M., Nácher, J., Varea, E., and Martínez-<br />

Guijarro, F. J. (2007). Migrat<strong>in</strong>g neuroblasts <strong>of</strong> the rostral migratory stream are putative<br />

targets for the action <strong>of</strong> nitric oxide. Eur. J. Neurosci 26, 392-402.<br />

Haase, A., and Bicker, G. (2003a). <strong>Nitric</strong> oxide and cyclic nucleotides are regulators <strong>of</strong> neuronal<br />

migration <strong>in</strong> an <strong>in</strong>sect embryo. Development 130, 3977-3987.<br />

Haase, A., and Bicker, G. (2003b). <strong>Nitric</strong> oxide and cyclic nucleotides are regulators <strong>of</strong> neuronal<br />

21


migration <strong>in</strong> an <strong>in</strong>sect embryo. Development 130, 3977-3987.<br />

Haghikia, A., Mergia, E., Friebe, A., Eysel, U. T., Koesl<strong>in</strong>g, D., and Mittmann, T. (2007). Longterm<br />

potentiation <strong>in</strong> the visual cortex requires both nitric oxide receptor guanylyl cyclases. J.<br />

Neurosci 27, 818-823.<br />

Hall, C.N., and Garthwaite J. (2009). What is the real physiological NO concentration <strong>in</strong> vivo?<br />

<strong>Nitric</strong> oxide 21, 92-103.<br />

Hartley, R. S., Margulis, M., Fishman, P. S., Lee, V. M., and Tang, C. M. (1999). Functional<br />

synapses are formed between human NTera2 (NT2N, hNT) neurons grown on astrocytes. J.<br />

Comp. Neurol 407, 1-10.<br />

Hatten, M. E. (1999). Central nervous system neuronal migration. Annu. Rev. Neurosci 22, 511-<br />

539.<br />

He, J., Wang, T., Wang, P., Han, P., Y<strong>in</strong>, Q., and Chen, C. (2007). A novel mechanism underly<strong>in</strong>g<br />

the susceptibility <strong>of</strong> neuronal cells to nitric oxide: the occurrence and regulation <strong>of</strong> prote<strong>in</strong><br />

S-nitrosylation is the checkpo<strong>in</strong>t. J. Neurochem 102, 1863-1874.<br />

He, Y., Yu, W., and Baas, P. W. (2002). Microtubule reconfiguration dur<strong>in</strong>g axonal retraction<br />

<strong>in</strong>duced by nitric oxide. J. Neurosci 22, 5982-5991.<br />

Hess, D. T., Patterson, S. I., Smith, D. S., and Skene, J. H. (1993). Neuronal growth cone collapse<br />

and <strong>in</strong>hibition <strong>of</strong> prote<strong>in</strong> fatty acylation by nitric oxide. Nature 366, 562-565.<br />

H<strong>in</strong>dley, S., Juurl<strong>in</strong>k, B. H., Gysbers, J. W., Middlemiss, P. J., Herman, M. A., and Rathbone, M. P.<br />

(1997). <strong>Nitric</strong> oxide donors enhance neurotroph<strong>in</strong>-<strong>in</strong>duced neurite outgrowth through a<br />

cGMP-dependent mechanism. J. Neurosci. Res 47, 427-439.<br />

Houldsworth, J., Heath, S. C., Bosl, G. J., Studer, L., and Chaganti, R. S. K. (2002). Expression<br />

pr<strong>of</strong>il<strong>in</strong>g <strong>of</strong> l<strong>in</strong>eage differentiation <strong>in</strong> pluripotential human embryonal carc<strong>in</strong>oma cells. Cell<br />

Growth Differ 13, 257-264.<br />

Ignarro, L. J., Byrns, R. E., Buga, G. M., and Wood, K. S. (1987). Endothelium-derived relax<strong>in</strong>g<br />

factor from pulmonary artery and ve<strong>in</strong> possesses pharmacologic and chemical properties<br />

identical to those <strong>of</strong> nitric oxide radical. Circ. Res 61, 866-879.<br />

J<strong>in</strong>, K., M<strong>in</strong>ami, M., Xie, L., Sun, Y., Mao, X. O., Wang, Y., Simon, R. P., and Greenberg, D. A.<br />

(2004a). Ischemia-<strong>in</strong>duced neurogenesis is preserved but reduced <strong>in</strong> the aged rodent bra<strong>in</strong>.<br />

Ag<strong>in</strong>g Cell 3, 373-377.<br />

J<strong>in</strong>, K., M<strong>in</strong>ami, M., Xie, L., Sun, Y., Mao, X. O., Wang, Y., Simon, R. P., and Greenberg, D. A.<br />

(2004b). Ischemia-<strong>in</strong>duced neurogenesis is preserved but reduced <strong>in</strong> the aged rodent bra<strong>in</strong>.<br />

Ag<strong>in</strong>g Cell 3, 373-377.<br />

Kater, S. B., and Rehder, V. (1995). The sensory-motor role <strong>of</strong> growth cone filopodia. Curr. Op<strong>in</strong>.<br />

Neurobiol 5, 68-74.<br />

Katsuki, S., Arnold, W., Mittal, C., and Murad, F. (1977). Stimulation <strong>of</strong> guanylate cyclase by<br />

sodium nitroprusside, nitroglycer<strong>in</strong> and nitric oxide <strong>in</strong> various tissue preparations and<br />

comparison to the effects <strong>of</strong> sodium azide and hydroxylam<strong>in</strong>e. J Cyclic Nucleotide Res 3,<br />

22


23-35.<br />

Knipp, S., and Bicker, G. (2009a). Regulation <strong>of</strong> enteric neuron migration by the gaseous<br />

messenger molecules CO and NO. Development 136, 85-93.<br />

Knipp, S., and Bicker, G. (2009b). Regulation <strong>of</strong> enteric neuron migration by the gaseous<br />

messenger molecules CO and NO. Development 136, 85-93.<br />

Knowles, R. G., and Moncada, S. (1994). <strong>Nitric</strong> oxide synthases <strong>in</strong> mammals. Biochem. J 298 ( Pt<br />

2), 249-258.<br />

Koriyama, Y., Yasuda, R., Homma, K., Mawatari, K., Nagashima, M., Sugitani, K., Matsukawa, T.,<br />

and Kato, S. (2009). <strong>Nitric</strong> oxide-cGMP signal<strong>in</strong>g regulates axonal elongation dur<strong>in</strong>g optic<br />

nerve regeneration <strong>in</strong> the goldfish <strong>in</strong> vitro and <strong>in</strong> vivo. J. Neurochem 110, 890-901.<br />

Kurosaki, R., Muramatsu, Y., Michimata, M., Matsubara, M., Kato, H., Imai, Y., Itoyama, Y., and<br />

Araki, T. (2002). Role <strong>of</strong> nitric oxide synthase aga<strong>in</strong>st MPTP neurotoxicity <strong>in</strong> mice. Neurol.<br />

Res 24, 655-662.<br />

Kuz<strong>in</strong>, B., Roberts, I., Peunova, N., and Enikolopov, G. (1996). <strong>Nitric</strong> oxide regulates cell<br />

proliferation dur<strong>in</strong>g Drosophila development. Cell 87, 639-649.<br />

L<strong>in</strong>dsay, S. L., Ramsey, S., Aitchison, M., Renné, T., and Evans, T. J. (2007). Modulation <strong>of</strong><br />

lamellipodial structure and dynamics by NO-dependent phosphorylation <strong>of</strong> VASP Ser239. J.<br />

Cell. Sci 120, 3011-3021.<br />

Luo, C. X., Zhu, X. J., Zhang, A. X., Wang, W., Yang, X. M., Liu, S. H., Han, X., Sun, J., Zhang,<br />

S. G., Lu, Y., et al. (2005). Blockade <strong>of</strong> L-type voltage-gated Ca channel <strong>in</strong>hibits ischemia<strong>in</strong>duced<br />

neurogenesis by down-regulat<strong>in</strong>g iNOS expression <strong>in</strong> adult mouse. J. Neurochem<br />

94, 1077-1086.<br />

Luo, C. X., Zhu, X. J., Zhou, Q. G., Wang, B., Wang, W., Cai, H. H., Sun, Y. J., Hu, M., Jiang, J.,<br />

Hua, Y., et al. (2007). Reduced neuronal nitric oxide synthase is <strong>in</strong>volved <strong>in</strong> ischemia<strong>in</strong>duced<br />

hippocampal neurogenesis by up-regulat<strong>in</strong>g <strong>in</strong>ducible nitric oxide synthase<br />

expression. J. Neurochem 103, 1872-1882.<br />

Lüth, H. J., Holzer, M., Gärtner, U., Staufenbiel, M., and Arendt, T. (2001). Expression <strong>of</strong><br />

endothelial and <strong>in</strong>ducible NOS-is<strong>of</strong>orms is <strong>in</strong>creased <strong>in</strong> Alzheimer's disease, <strong>in</strong> APP23<br />

transgenic mice and after experimental bra<strong>in</strong> lesion <strong>in</strong> rat: evidence for an <strong>in</strong>duction by<br />

amyloid pathology. Bra<strong>in</strong> Res 913, 57-67.<br />

Madhusoodanan, K. S., and Murad, F. (2007). NO-cGMP signal<strong>in</strong>g and regenerative medic<strong>in</strong>e<br />

<strong>in</strong>volv<strong>in</strong>g stem cells. Neurochem. Res 32, 681-694.<br />

Marín, O., and Rubenste<strong>in</strong>, J. L. R. (2003). Cell migration <strong>in</strong> the forebra<strong>in</strong>. Annu. Rev. Neurosci 26,<br />

441-483.<br />

Matteoli, M., Coco, S., Schenk, U., and Verderio, C. (2004). Vesicle turnover <strong>in</strong> develop<strong>in</strong>g<br />

neurons: how to build a presynaptic term<strong>in</strong>al. Trends Cell Biol 14, 133-140.<br />

Matthews, R. T., Beal, M. F., Fallon, J., Fedorchak, K., Huang, P. L., Fishman, M. C., and Hyman,<br />

B. T. (1997). MPP+ <strong>in</strong>duced substantia nigra degeneration is attenuated <strong>in</strong> nNOS knockout<br />

mice. Neurobiol. Dis 4, 114-121.<br />

23


McAllister, A. K. (2007). Dynamic aspects <strong>of</strong> CNS synapse formation. Annu. Rev. Neurosci 30,<br />

425-450.<br />

Menegon, A., Bonanomi, D., Albert<strong>in</strong>azzi, C., Lotti, F., Ferrari, G., Kao, H., Benfenati, F., Baldelli,<br />

P., and Valtorta, F. (2006). Prote<strong>in</strong> k<strong>in</strong>ase A-mediated synaps<strong>in</strong> I phosphorylation is a<br />

central modulator <strong>of</strong> Ca2+-dependent synaptic activity. J. Neurosci 26, 11670-11681.<br />

Mét<strong>in</strong>, C., Vallee, R. B., Rakic, P., and Bhide, P. G. (2008). Modes and mishaps <strong>of</strong> neuronal<br />

migration <strong>in</strong> the mammalian bra<strong>in</strong>. J. Neurosci 28, 11746-11752.<br />

Micheva, K. D., Buchanan, J., Holz, R. W., and Smith, S. J. (2003). Retrograde regulation <strong>of</strong><br />

synaptic vesicle endocytosis and recycl<strong>in</strong>g. Nat. Neurosci 6, 925-932.<br />

Moors, M., Cl<strong>in</strong>e, J. E., Abel, J., and Fritsche, E. (2007). ERK-dependent and -<strong>in</strong>dependent<br />

pathways trigger human neural progenitor cell migration. Toxicol. Appl. Pharmacol 221, 57-<br />

67.<br />

Moors, M., Rockel, T. D., Abel, J., Cl<strong>in</strong>e, J. E., Gassmann, K., Schreiber, T., Schuwald, J.,<br />

We<strong>in</strong>mann, N., and Fritsche, E. (2009). <strong>Human</strong> neurospheres as three-dimensional cellular<br />

systems for developmental neurotoxicity test<strong>in</strong>g. Environ. Health Perspect 117, 1131-1138.<br />

Moreno-López, B., Noval, J. A., González-Bonet, L. G., and Estrada, C. (2000). Morphological<br />

bases for a role <strong>of</strong> nitric oxide <strong>in</strong> adult neurogenesis. Bra<strong>in</strong> Res 869, 244-250.<br />

Moreno-López, B., Romero-Grimaldi, C., Noval, J. A., Murillo-Carretero, M., Matarredona, E. R.,<br />

and Estrada, C. (2004). <strong>Nitric</strong> oxide is a physiological <strong>in</strong>hibitor <strong>of</strong> neurogenesis <strong>in</strong> the adult<br />

mouse subventricular zone and olfactory bulb. J. Neurosci 24, 85-95.<br />

Murad, F. (1986). Cyclic guanos<strong>in</strong>e monophosphate as a mediator <strong>of</strong> vasodilation. J. Cl<strong>in</strong>. Invest<br />

78, 1-5.<br />

Murillo-Carretero, M., Ruano, M. J., Matarredona, E. R., Villalobo, A., and Estrada, C. (2002).<br />

Antiproliferative effect <strong>of</strong> nitric oxide on epidermal growth factor-responsive human<br />

neuroblastoma cells. J. Neurochem 83, 119-131.<br />

Müller, U. (1997). The nitric oxide systems <strong>in</strong> <strong>in</strong>sects. Prog. Neurobiol. 51, 363-381.<br />

Nikonenko, I., Jourda<strong>in</strong>, P., Muller, D. (2003). Presynaptic remodel<strong>in</strong>g contributes to activitydependent<br />

synaptogenesis. J. Neurosci. 23, 8498-8505.<br />

Nikonenko, I., Boda, B., Steen, S., Knott, G., Welker, E., and Muller, D. (2008). PSD-95 promotes<br />

synaptogenesis and multi<strong>in</strong>nervated sp<strong>in</strong>e formation through nitric oxide signal<strong>in</strong>g. J. Cell<br />

Biol 183, 1115-1127.<br />

Nott, A., and Riccio, A. (2009). <strong>Nitric</strong> oxide-mediated epigenetic mechanisms <strong>in</strong> develop<strong>in</strong>g<br />

neurons. Cell Cycle 8, 725-730.<br />

Nott, A., Watson, P. M., Rob<strong>in</strong>son, J. D., Crepaldi, L., and Riccio, A. (2008). S-Nitrosylation <strong>of</strong><br />

histone deacetylase 2 <strong>in</strong>duces chromat<strong>in</strong> remodell<strong>in</strong>g <strong>in</strong> neurons. Nature 455, 411-415.<br />

Nott, A., Veenvliet, J., Goda, Y., Riccio, A.(2009). S-Nitrosylation <strong>of</strong> the chromat<strong>in</strong>–remodell<strong>in</strong>g<br />

prote<strong>in</strong> histone deacetylase 2 regulates cortical development. Program NO. 219.14. 2009<br />

Neuroscience Meet<strong>in</strong>g Planner. Chicago, IL: Society for Neuroscience meet<strong>in</strong>g,<br />

2009.Onl<strong>in</strong>e.<br />

24


Ogilvie, P., Schill<strong>in</strong>g, K., Bill<strong>in</strong>gsley, M. L., and Schmidt, H. H. (1995). Induction and variants <strong>of</strong><br />

neuronal nitric oxide synthase type I dur<strong>in</strong>g synaptogenesis. FASEB J 9, 799-806.<br />

Packer, M. A., Stasiv, Y., Benraiss, A., Chmielnicki, E., Gr<strong>in</strong>berg, A., Westphal, H., Goldman, S.<br />

A., and Enikolopov, G. (2003). <strong>Nitric</strong> oxide negatively regulates mammalian adult<br />

neurogenesis. Proc. Natl. Acad. Sci. U.S.A 100, 9566-9571.<br />

Palmer, R. M., Ferrige, A. G., and Moncada, S. (1987). <strong>Nitric</strong> oxide release accounts for the<br />

biological activity <strong>of</strong> endothelium-derived relax<strong>in</strong>g factor. Nature 327, 524-526.<br />

Paquet-Durand, F., and Bicker, G. (2007). <strong>Human</strong> model neurons <strong>in</strong> studies <strong>of</strong> bra<strong>in</strong> cell damage<br />

and neural repair. Curr. Mol. Med 7, 541-554.<br />

Paquet-Durand, F., Tan, S., and Bicker, G. (2003). Turn<strong>in</strong>g teratocarc<strong>in</strong>oma cells <strong>in</strong>to neurons:<br />

rapid differentiation <strong>of</strong> NT-2 cells <strong>in</strong> float<strong>in</strong>g spheres. Bra<strong>in</strong> Res. Dev. Bra<strong>in</strong> Res 142, 161-<br />

167.<br />

Park, C., Sohn, Y., Sh<strong>in</strong>, K. S., Kim, J., Ahn, H., and Huh, Y. (2003). The chronic <strong>in</strong>hibition <strong>of</strong><br />

nitric oxide synthase enhances cell proliferation <strong>in</strong> the adult rat hippocampus. Neurosci. Lett<br />

339, 9-12.<br />

Petrov, A. M., G<strong>in</strong>iatull<strong>in</strong>, A. R., Sitdikova, G. F., and Zefirov, A. L. (2008). The role <strong>of</strong> cGMPdependent<br />

signal<strong>in</strong>g pathway <strong>in</strong> synaptic vesicle cycle at the frog motor nerve term<strong>in</strong>als. J.<br />

Neurosci 28, 13216-13222.<br />

Peunova, N., and Enikolopov, G. (1995). <strong>Nitric</strong> oxide triggers a switch to growth arrest dur<strong>in</strong>g<br />

differentiation <strong>of</strong> neuronal cells. Nature 375, 68-73.<br />

Peunova, N., Sche<strong>in</strong>ker, V., Cl<strong>in</strong>e, H., and Enikolopov, G. (2001). <strong>Nitric</strong> oxide is an essential<br />

negative regulator <strong>of</strong> cell proliferation <strong>in</strong> Xenopus bra<strong>in</strong>. J. Neurosci 21, 8809-8818.<br />

Peunova, N., Sche<strong>in</strong>ker, V., Ravi, K., and Enikolopov, G. (2007). <strong>Nitric</strong> oxide coord<strong>in</strong>ates cell<br />

proliferation and cell movements dur<strong>in</strong>g early development <strong>of</strong> Xenopus. Cell Cycle 6, 3132-<br />

3144.<br />

Phung, Y. T., Bekker, J. M., Hallmark, O. G., and Black, S. M. (1999). Both neuronal NO synthase<br />

and nitric oxide are required for PC12 cell differentiation: a cGMP <strong>in</strong>dependent pathway.<br />

Bra<strong>in</strong> Res. Mol. Bra<strong>in</strong> Res 64, 165-178.<br />

Pleasure, S. J., Page, C., and Lee, V. M. (1992). Pure, postmitotic, polarized human neurons derived<br />

from NTera 2 cells provide a system for express<strong>in</strong>g exogenous prote<strong>in</strong>s <strong>in</strong> term<strong>in</strong>ally<br />

differentiated neurons. J. Neurosci 12, 1802-1815.<br />

Podrygajlo, G., Tegenge, M. A., Gierse, A., Paquet-Durand, F., Tan, S., Bicker, G., and Stern, M.<br />

(2009). Cellular phenotypes <strong>of</strong> human model neurons (NT2) after differentiation <strong>in</strong><br />

aggregate culture. Cell Tissue Res 336, 439-452.<br />

Podrygajlo, G, Song, Y, Schles<strong>in</strong>ger, F, Krampfl, K, Bicker G.(2010). Synaptic currents and<br />

transmitter responses <strong>in</strong> human NT2 neurons differentiated <strong>in</strong> aggregate culture. Neurosci. Lett (In<br />

press).<br />

Przedborski, S., Jackson-Lewis, V., Yokoyama, R., Shibata, T., Dawson, V. L., and Dawson, T. M.<br />

(1996). Role <strong>of</strong> neuronal nitric oxide <strong>in</strong> 1-methyl-4-phenyl-1,2,3,6-tetrahydropyrid<strong>in</strong>e<br />

25


(MPTP)-<strong>in</strong>duced dopam<strong>in</strong>ergic neurotoxicity. Proc. Natl. Acad. Sci. U.S.A 93, 4565-4571.<br />

Przyborski, S. A., Christie, V. B., Hayman, M. W., Stewart, R., and Horrocks, G. M. (2004).<br />

<strong>Human</strong> embryonal carc<strong>in</strong>oma stem cells: models <strong>of</strong> embryonic development <strong>in</strong> humans.<br />

Stem Cells Dev 13, 400-408.<br />

Przyborski, S. A., Morton, I. E., Wood, A., and Andrews, P. W. (2000). Developmental regulation<br />

<strong>of</strong> neurogenesis <strong>in</strong> the pluripotent human embryonal carc<strong>in</strong>oma cell l<strong>in</strong>e NTERA-2. Eur. J.<br />

Neurosci 12, 3521-3528.<br />

Przyborski, S. A., Smith, S., and Wood, A. (2003). Transcriptional pr<strong>of</strong>il<strong>in</strong>g <strong>of</strong> neuronal<br />

differentiation by human embryonal carc<strong>in</strong>oma stem cells <strong>in</strong> vitro. Stem Cells 21, 459-471.<br />

Reif, A., Schmitt, A., Fritzen, S., Chourbaji, S., Bartsch, C., Urani, A., Wycislo, M., Mössner, R.,<br />

Sommer, C., Gass, P., et al. (2004). Differential effect <strong>of</strong> endothelial nitric oxide synthase<br />

(NOS-III) on the regulation <strong>of</strong> adult neurogenesis and behaviour. Eur. J. Neurosci 20, 885-<br />

895.<br />

Rentería, R. C., and Constant<strong>in</strong>e-Paton, M. (1996). Exogenous nitric oxide causes collapse <strong>of</strong> ret<strong>in</strong>al<br />

ganglion cell axonal growth cones <strong>in</strong> vitro. J. Neurobiol 29, 415-428.<br />

Rettori, V., Fernandez-Solari, J., Mohn, C., Zorrilla Zubilete, M. A., de la Cal, C., Prestifilippo, J.<br />

P., and De Laurentiis, A. (2009). <strong>Nitric</strong> oxide at the crossroad <strong>of</strong> immunoneuroendocr<strong>in</strong>e<br />

<strong>in</strong>teractions. Ann. N. Y. Acad. Sci 1153, 35-47.<br />

Rialas, C. M., Nomizu, M., Patterson, M., Kle<strong>in</strong>man, H. K., Weston, C. A., and Weeks, B. S.<br />

(2000). <strong>Nitric</strong> oxide mediates lam<strong>in</strong><strong>in</strong>-<strong>in</strong>duced neurite outgrowth <strong>in</strong> PC12 cells. Exp. Cell<br />

Res 260, 268-276.<br />

Riccio, A., Alvania, R. S., Lonze, B. E., Ramanan, N., Kim, T., Huang, Y., Dawson, T. M., Snyder,<br />

S. H., and G<strong>in</strong>ty, D. D. (2006). A nitric oxide signal<strong>in</strong>g pathway controls CREB-mediated<br />

gene expression <strong>in</strong> neurons. Mol. Cell 21, 283-294.<br />

Santacana, M., Uttenthal, L. O., Bentura, M. L., Fernández, A. P., Serrano, J., Martínez de Velasco,<br />

J., Alonso, D., Martínez-Murillo, R., and Rodrigo, J. (1998). Expression <strong>of</strong> neuronal nitric<br />

oxide synthase dur<strong>in</strong>g embryonic development <strong>of</strong> the rat cerebral cortex. Bra<strong>in</strong> Res. Dev.<br />

Bra<strong>in</strong> Res 111, 205-222.<br />

Schonh<strong>of</strong>f, C. M., Bulseco, D. A., Brancho, D. M., Parada, L. F., and Ross, A. H. (2001). The Ras-<br />

ERK pathway is required for the <strong>in</strong>duction <strong>of</strong> neuronal nitric oxide synthase <strong>in</strong><br />

differentiat<strong>in</strong>g PC12 cells. J. Neurochem 78, 631-639.<br />

Seidel, C., and Bicker, G. (2000). <strong>Nitric</strong> oxide and cGMP <strong>in</strong>fluence axonogenesis <strong>of</strong> antennal<br />

pioneer neurons. Development 127, 4541-4549.<br />

Seidel, C., and Bicker, G. (2002). Developmental expression <strong>of</strong> nitric oxide/cyclic GMP signal<strong>in</strong>g<br />

pathways <strong>in</strong> the bra<strong>in</strong> <strong>of</strong> the embryonic grasshopper. Bra<strong>in</strong> Res. Dev. Bra<strong>in</strong> Res 138, 71-79.<br />

Shan, X., Chi, L., Bishop, M., Luo, C., Lien, L., Zhang, Z., and Liu, R. (2006). Enhanced de novo<br />

neurogenesis and dopam<strong>in</strong>ergic neurogenesis <strong>in</strong> the substantia nigra <strong>of</strong> 1-methyl-4-phenyl-<br />

1,2,3,6-tetrahydropyrid<strong>in</strong>e-<strong>in</strong>duced Park<strong>in</strong>son's disease-like mice. Stem Cells 24, 1280-<br />

1287.<br />

26


Silverman, R. B. (2009). Design <strong>of</strong> Selective Neuronal <strong>Nitric</strong> <strong>Oxide</strong> Synthase Inhibitors for the<br />

Prevention and Treatment <strong>of</strong> Neurodegenerative Diseases. Acc. Chem. Res. Available at:<br />

http://www.ncbi.nlm.nih.gov/pubmed/19154146 [Accessed October 7, 2009].<br />

Simic, G., Lucassen, P. J., Krsnik, Z., Krusl<strong>in</strong>, B., Kostovic, I., W<strong>in</strong>blad, B., and Bogdanovi (2000).<br />

nNOS expression <strong>in</strong> reactive astrocytes correlates with <strong>in</strong>creased cell death related DNA<br />

damage <strong>in</strong> the hippocampus and entorh<strong>in</strong>al cortex <strong>in</strong> Alzheimer's disease. Exp. Neurol 165,<br />

12-26.<br />

Simmons, M. L., and Murphy, S. (1992). Induction <strong>of</strong> nitric oxide synthase <strong>in</strong> glial cells. J.<br />

Neurochem 59, 897-905.<br />

Spitzer, N. C. (2006). Electrical activity <strong>in</strong> early neuronal development. Nature 444, 707-712.<br />

Sporbert, A., Mertsch, K., Smolenski, A., Hasel<strong>of</strong>f, R. F., Schönfelder, G., Paul, M., Ruth, P.,<br />

Walter, U., and Blasig, I. E. (1999). Phosphorylation <strong>of</strong> vasodilator-stimulated<br />

phosphoprote<strong>in</strong>: a consequence <strong>of</strong> nitric oxide- and cGMP-mediated signal transduction <strong>in</strong><br />

bra<strong>in</strong> capillary endothelial cells and astrocytes. Bra<strong>in</strong> Res. Mol. Bra<strong>in</strong> Res 67, 258-266.<br />

Sporns, O., and Jenk<strong>in</strong>son, S. (1997). Potassium ion- and nitric oxide-<strong>in</strong>duced exocytosis from<br />

populations <strong>of</strong> hippocampal synapses dur<strong>in</strong>g synaptic maturation <strong>in</strong> vitro. Neuroscience 80,<br />

1057-1073.<br />

Stern, M., and Bicker, G. (2008). <strong>Nitric</strong> oxide regulates axonal regeneration <strong>in</strong> an <strong>in</strong>sect embryonic<br />

CNS. Dev Neurobiol 68, 295-308.<br />

Stroissnigg, H., Trancíková, A., Descovich, L., Fuhrmann, J., Kutschera, W., Kostan, J., Meixner,<br />

A., Nothias, F., and Propst, F. (2007). S-nitrosylation <strong>of</strong> microtubule-associated prote<strong>in</strong> 1B<br />

mediates nitric-oxide-<strong>in</strong>duced axon retraction. Nat. Cell Biol 9, 1035-1045.<br />

Sultana, R., Poon, H. F., Cai, J., Pierce, W. M., Merchant, M., Kle<strong>in</strong>, J. B., Markesbery, W. R., and<br />

Butterfield, D. A. (2006). Identification <strong>of</strong> nitrated prote<strong>in</strong>s <strong>in</strong> Alzheimer's disease bra<strong>in</strong><br />

us<strong>in</strong>g a redox proteomics approach. Neurobiol. Dis 22, 76-87.<br />

Sun, Y., J<strong>in</strong>, K., Childs, J. T., Xie, L., Mao, X. O., and Greenberg, D. A. (2005). Neuronal nitric<br />

oxide synthase and ischemia-<strong>in</strong>duced neurogenesis. J. Cereb. Blood Flow Metab 25, 485-<br />

492.<br />

Schwartz, C.M., Spivak, C.E., Baker, S.C., McDaniel, T.K., Lor<strong>in</strong>g, J.F., Nguyen, C., Chrest, F.J.,<br />

Wersto, R., Arenas, E., Zeng, X., Freed, W.J., Rao, M.S. (2005). NTera2: a model system to<br />

study dopam<strong>in</strong>ergic differentiation <strong>of</strong> human embryonic stem cells. Stem cells Dev. 14, 517-<br />

534.<br />

Tanaka, M., Yoshida, S., Yano, M., and Hanaoka, F. (1994). Roles <strong>of</strong> endogenous nitric oxide <strong>in</strong><br />

cerebellar cortical development <strong>in</strong> slice cultures. Neuroreport 5, 2049-2052.<br />

Tegenge, M. A., and Bicker, G. (2009). <strong>Nitric</strong> oxide and cGMP signal transduction positively<br />

regulates the motility <strong>of</strong> human neuronal precursor (NT2) cells. J. Neurochem 110, 1828-<br />

1841.<br />

Tegenge, M. A., Stern, M., and Bicker, G. (2009). <strong>Nitric</strong> oxide and cyclic nucleotide signal<br />

transduction modulates synaptic vesicle turnover <strong>in</strong> human model neurons. J. Neurochem.<br />

27


111, 1434-1446.<br />

Thippeswamy, T., and Morris, R. (2002). The roles <strong>of</strong> nitric oxide <strong>in</strong> dorsal root ganglion neurons.<br />

Ann. N. Y. Acad. Sci 962, 103-110.<br />

Thorns, V., Hansen, L., and Masliah, E. (1998). nNOS express<strong>in</strong>g neurons <strong>in</strong> the entorh<strong>in</strong>al cortex<br />

and hippocampus are affected <strong>in</strong> patients with Alzheimer's disease. Exp. Neurol 150, 14-20.<br />

Togashi, K., von Schimmelmann, M. J., Nishiyama, M., Lim, C., Yoshida, N., Yun, B., Molday, R.<br />

S., Goshima, Y., and Hong, K. (2008). Cyclic GMP-gated CNG channels function <strong>in</strong><br />

Sema3A-<strong>in</strong>duced growth cone repulsion. Neuron 58, 694-707.<br />

Tojima, T., It<strong>of</strong>usa, R., and Kamiguchi, H. (2009). The nitric oxide-cGMP pathway controls the<br />

directional polarity <strong>of</strong> growth cone guidance via modulat<strong>in</strong>g cytosolic Ca2+ signals. J.<br />

Neurosci 29, 7886-7897.<br />

Traister, A., Abashidze, S., Gold, V., Plachta, N., Karchovsky, E., Patel, K., and Weil, M. (2002).<br />

Evidence that nitric oxide regulates cell-cycle progression <strong>in</strong> the develop<strong>in</strong>g chick<br />

neuroepithelium. Dev. Dyn 225, 271-276.<br />

Trimm, K. R., and Rehder, V. (2004). <strong>Nitric</strong> oxide acts as a slow-down and search signal <strong>in</strong><br />

develop<strong>in</strong>g neurites. Eur. J. Neurosci 19, 809-818.<br />

Truman, J. W., De Vente, J., and Ball, E. E. (1996). <strong>Nitric</strong> oxide-sensitive guanylate cyclase activity<br />

is associated with the maturational phase <strong>of</strong> neuronal development <strong>in</strong> <strong>in</strong>sects. Development<br />

122, 3949-3958.<br />

Waites, C. L., Craig, A. M., and Garner, C. C. (2005). Mechanisms <strong>of</strong> vertebrate synaptogenesis.<br />

Annu. Rev. Neurosci 28, 251-274.<br />

Watanabe, Y., Kato, H., and Araki, T. (2008). Protective action <strong>of</strong> neuronal nitric oxide synthase<br />

<strong>in</strong>hibitor <strong>in</strong> the MPTP mouse model <strong>of</strong> Park<strong>in</strong>son's disease. Metab Bra<strong>in</strong> Dis 23, 51-69.<br />

Welshhans, K., and Rehder, V. (2005). Local activation <strong>of</strong> the nitric oxide/cyclic guanos<strong>in</strong>e<br />

monophosphate pathway <strong>in</strong> growth cones regulates filopodial length via prote<strong>in</strong> k<strong>in</strong>ase G,<br />

cyclic ADP ribose and <strong>in</strong>tracellular Ca2+ release. Eur. J. Neurosci 22, 3006-3016.<br />

Wernig, M., Zhao, J., Pruszak, J., Hedlund, E., Fu, D., Soldner, F., Broccoli, V., Constant<strong>in</strong>e-Paton,<br />

M., Isacson, O., and Jaenisch, R. (2008). Neurons derived from reprogrammed fibroblasts<br />

functionally <strong>in</strong>tegrate <strong>in</strong>to the fetal bra<strong>in</strong> and improve symptoms <strong>of</strong> rats with Park<strong>in</strong>son's<br />

disease. Proc. Natl. Acad. Sci. U.S.A 105, 5856-5861.<br />

Wessely, O., and De Robertis, E. M. (2002). Neural plate pattern<strong>in</strong>g by secreted signals. Neuron 33,<br />

489-491.<br />

Wester<strong>in</strong>k, R. H. S., and Ew<strong>in</strong>g, A. G. (2008). The PC12 cell as model for neurosecretion. Acta<br />

Physiol (Oxf) 192, 273-285.<br />

Wildemann, B., and Bicker, G. (1999). Developmental expression <strong>of</strong> nitric oxide/cyclic GMP<br />

synthesiz<strong>in</strong>g cells <strong>in</strong> the nervous system <strong>of</strong> Drosophila melanogaster. J. Neurobiol 38, 1-15.<br />

Williams, C. V., Nordquist, D., and McLoon, S. C. (1994). Correlation <strong>of</strong> nitric oxide synthase<br />

expression with chang<strong>in</strong>g patterns <strong>of</strong> axonal projections <strong>in</strong> the develop<strong>in</strong>g visual system. J.<br />

Neurosci 14, 1746-1755.<br />

28


Yamazaki, M., Chiba, K., Mohri, T., and Hatanaka, H. (2001). Activation <strong>of</strong> the mitogen-activated<br />

prote<strong>in</strong> k<strong>in</strong>ase cascade through nitric oxide synthesis as a mechanism <strong>of</strong> neuritogenic effect<br />

<strong>of</strong> genip<strong>in</strong> <strong>in</strong> PC12h cells. J. Neurochem 79, 45-54.<br />

Yoshihara, Y., De Roo, M., and Muller, D. (2009). Dendritic sp<strong>in</strong>e formation and stabilization.<br />

Curr. Op<strong>in</strong>. Neurobiol 19, 146-153.<br />

Young, D. V., Serebryanik, D., Janero, D. R., and Tam, S. W. (2000). Suppression <strong>of</strong> proliferation<br />

<strong>of</strong> human coronary artery smooth muscle cells by the nitric oxide donor, Snitrosoglutathione,<br />

is cGMP-<strong>in</strong>dependent. Mol. Cell Biol. Res. Commun 4, 32-36.<br />

Yue, X., Dreyfus, C., Kong, T. A., and Zhou, R. (2008). A subset <strong>of</strong> signal transduction pathways is<br />

required for hippocampal growth cone collapse <strong>in</strong>duced by ephr<strong>in</strong>-A5. Dev Neurobiol 68,<br />

1269-1286.<br />

Zhang, P., Liu, Y., Li, J., Kang, Q., Tian, Y., Chen, X., Zhao, J., Shi, Q., and Song, T. (2007).<br />

Decreased neuronal nitric oxide synthase expression and cell migration <strong>in</strong> the peri-<strong>in</strong>farction<br />

after focal cerebral ischemia <strong>in</strong> rats. Neuropathology 27, 347-354.<br />

Zhang, R., Zhang, L., Zhang, Z., Wang, Y., Lu, M., Lapo<strong>in</strong>te, M., and Chopp, M. (2001). A nitric<br />

oxide donor <strong>in</strong>duces neurogenesis and reduces functional deficits after stroke <strong>in</strong> rats. Ann.<br />

Neurol 50, 602-611.<br />

Zhou, L., and Zhu, D. (2009a). Neuronal nitric oxide synthase: structure, subcellular localization,<br />

regulation, and cl<strong>in</strong>ical implications. <strong>Nitric</strong> <strong>Oxide</strong> 20, 223-230.<br />

Zhou, L., and Zhu, D. (2009b). Neuronal nitric oxide synthase: structure, subcellular localization,<br />

regulation, and cl<strong>in</strong>ical implications. <strong>Nitric</strong> <strong>Oxide</strong> 20, 223-230.<br />

Zhu, D. Y., Liu, S. H., Sun, H. S., and Lu, Y. M. (2003). Expression <strong>of</strong> <strong>in</strong>ducible nitric oxide<br />

synthase after focal cerebral ischemia stimulates neurogenesis <strong>in</strong> the adult rodent dentate<br />

gyrus. J. Neurosci 23, 223-229.<br />

Zhu, X. J., Hua, Y., Jiang, J., Zhou, Q. G., Luo, C. X., Han, X., Lu, Y. M., and Zhu, D. Y. (2006).<br />

Neuronal nitric oxide synthase-derived nitric oxide <strong>in</strong>hibits neurogenesis <strong>in</strong> the adult dentate<br />

gyrus by down-regulat<strong>in</strong>g cyclic AMP response element b<strong>in</strong>d<strong>in</strong>g prote<strong>in</strong> phosphorylation.<br />

Neuroscience 141, 827-836.<br />

Zulauf, L., Coste, O., Marian, C., Möser, C., Brenneis, C., Niederberger, E. (2009). C<strong>of</strong>il<strong>in</strong><br />

phosphorylation is <strong>in</strong>volved <strong>in</strong> nitric oxide/cGMP-mediated nociception. Biochem. Biophys.<br />

Res. Commun 390, 1408-1413.<br />

29


Acknowledgments<br />

I thank my supervisor Pr<strong>of</strong>. Dr. Gerd Bicker for allow<strong>in</strong>g me to work <strong>in</strong> an excit<strong>in</strong>g project and for<br />

his excellent guidance throughout my PhD project. Dur<strong>in</strong>g my stay <strong>in</strong> your laboratory I have<br />

learned not only how to become a good researcher but also a mentor. Thank you very much!<br />

I also extend my s<strong>in</strong>cere gratitude to my advisory group Pr<strong>of</strong>. Dr. Stephan Ste<strong>in</strong>lechner and Pr<strong>of</strong>.<br />

Dr. Herbert Hildebrandt for the useful discussion, suggestion and guidance throughout my doctoral<br />

process.<br />

I thank Pr<strong>of</strong>. Dr.med. Ellen Fritsche and Dr. Thomas D<strong>in</strong>o Rockel (Group <strong>of</strong> Molecular Toxicology,<br />

Institut für Umweltmediz<strong>in</strong>ische Forschung, He<strong>in</strong>rich He<strong>in</strong>e-University, Dusseldorf, Germany) for<br />

permitt<strong>in</strong>g me to work on fetal human neural progenitor cells and cont<strong>in</strong>uous collaboration<br />

throughout this project.<br />

I thank ZSN PhD commission specially Pr<strong>of</strong>. Dr. Wolfgang Löscher, Pr<strong>of</strong>. Dr. Wolfgang<br />

Baumgartner and the coord<strong>in</strong>ation <strong>of</strong>fice (Dr. Dagmar Esser, Nadja Borsum and Manuela<br />

Chambers) for cont<strong>in</strong>uous support throught my doctoral study. I also s<strong>in</strong>cerely thank the M<strong>in</strong>istry<br />

for Science and Culture <strong>of</strong> Lower Saxony for giv<strong>in</strong>g me the Georg-Christoph-Lichtenberg<br />

scholarship throughout my PhD study.<br />

My stay <strong>in</strong> cell biology group could not be as fruitful as it is without the support and pleasant<br />

environment created by present and past colleagues <strong>in</strong> the lab. I thank Dr. Micheal Stern for his<br />

constructive comments, suggestion and discussion. I thank Saime Tan for her excellent technical<br />

support. I am also thankful to Dr. Sab<strong>in</strong>e Knipp and Dr. Grzegorz Podrygajlo for useful discussion.<br />

My thanks also go to Arne Pätschke, Andrea Gierse, Rene Eickh<strong>of</strong>f, Nicole Böger, Frank Rol<strong>of</strong>f<br />

and Christ<strong>in</strong>a Lorbeer.<br />

Last but not least I thank my family; Lydia and Betania. Thank you very much Lydia for the love,<br />

courage and support you gave me throughout this work. Betu you helped a lot by stay<strong>in</strong>g long <strong>in</strong> the<br />

k<strong>in</strong>dergarten. Thank you for your patience.<br />

30


Summary<br />

The complex mammalian central nervous system (CNS) undergoes sequential but partially<br />

overlapp<strong>in</strong>g cellular developmental processes that <strong>in</strong>clude cell proliferation, migration,<br />

differentiation and synaptogenesis. The gaseous messenger nitric oxide (NO) has been implicated to<br />

regulate each <strong>of</strong> these early neuronal developmental processes. However, it is not known whether<br />

NO signal<strong>in</strong>g is <strong>in</strong>volved <strong>in</strong> the development <strong>of</strong> human CNS. Here, I used human neuronal stem<br />

cells as a model <strong>of</strong> develop<strong>in</strong>g nerve cells to <strong>in</strong>vestigate the potential role <strong>of</strong> NO signal<strong>in</strong>g <strong>in</strong><br />

neuronal migration, differentiation and pre-synaptic vesicle release.<br />

In the first part, a model <strong>of</strong> human neuronal precursor cells was established from embryonic<br />

carc<strong>in</strong>oma stem cells (NT2 cells) as three-dimensional free-float<strong>in</strong>g aggregate culture. Upon<br />

ret<strong>in</strong>oic acid treatment the NT2 spherical aggregates generate cells that proliferate, migrate and<br />

differentiate <strong>in</strong>to neuronal phenotypes which mimic cellular processes that occur dur<strong>in</strong>g early<br />

development <strong>of</strong> CNS. Interest<strong>in</strong>gly, the migrat<strong>in</strong>g cells and cells <strong>in</strong>side the spheres respond to<br />

exogenous stimulation with NO by synthesiz<strong>in</strong>g cyclic guanos<strong>in</strong>e-monophosphate (cGMP),<br />

<strong>in</strong>dicat<strong>in</strong>g that the NT2 spherical aggregates express functional NO/cGMP signal<strong>in</strong>g. By employ<strong>in</strong>g<br />

small bioactive enzyme activators and <strong>in</strong>hibitors <strong>of</strong> neuronal nitric oxide synthase (nNOS), soluble<br />

guanylyl cyclase (sGC) and prote<strong>in</strong> k<strong>in</strong>ase G (PKG), I showed that the NO/cGMP/PKG signal<strong>in</strong>g<br />

cascade is a positive regulator <strong>of</strong> the migration <strong>of</strong> human neuronal precursor cells. Moreover, antiproliferative<br />

effect <strong>of</strong> NO was observed at relatively higher concentration <strong>of</strong> a NO donor. A modest<br />

but none significant effect <strong>of</strong> NO/cGMP signal<strong>in</strong>g on neuronal differentiation was also noted.<br />

In the second part, pre-synaptic maturation <strong>of</strong> human NT2 neurons were followed after the cells<br />

migrate out <strong>of</strong> NT2 spherical aggregates and form neuronal networks. <strong>Human</strong> NT2 neurons express<br />

<strong>in</strong>creas<strong>in</strong>g levels <strong>of</strong> the axonal marker (Tau) and typical pre-synaptic prote<strong>in</strong>s (synaps<strong>in</strong> and<br />

synaptotagm<strong>in</strong> I) depend<strong>in</strong>g on the length <strong>of</strong> <strong>in</strong> vitro culture. By employ<strong>in</strong>g an antibody directed<br />

aga<strong>in</strong>st the lum<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> synaptotagm<strong>in</strong> I and the fluorescent dye (FM1-43), I showed that<br />

depolarized mature NT2 neurons display calcium-dependent exo-endocytotic synaptic vesicle<br />

recycl<strong>in</strong>g. NT2 neurons express nNOS and a functional NO receptor, sGC. Next, I exam<strong>in</strong>ed<br />

whether NO signal transduction modulates synaptic vesicle turnover, a pathway that may contribute<br />

to activity dependent synaptic maturation <strong>in</strong> human model neurons. Two NO donors (sodium<br />

nitroprusside, and N-Ethyl-2-(1-ethyl-2-hydroxy-2-nitrosohydraz<strong>in</strong>o ethanam<strong>in</strong>e)) and cGMP<br />

analog (8-Br-cGMP) enhanced synaptic vesicle turnover while a sGC <strong>in</strong>hibitor blocked the effect <strong>of</strong><br />

NO donors. NO donors evoked vesicle exocytosis which was partially blocked by the sGC<br />

<strong>in</strong>hibitor. The activator <strong>of</strong> adenylyl cyclase, forskol<strong>in</strong>, and a cAMP analog <strong>in</strong>duced synaptic vesicle<br />

recycl<strong>in</strong>g and exocytosis via a parallel act<strong>in</strong>g prote<strong>in</strong> k<strong>in</strong>ase A pathway. These data suggest that<br />

31


NO/cyclic nucleotide signal<strong>in</strong>g pathways may facilitate pre-synaptic neurotransmitter release <strong>in</strong><br />

human bra<strong>in</strong>.<br />

F<strong>in</strong>ally, I tested the presence and function <strong>of</strong> NO/cGMP signal<strong>in</strong>g <strong>in</strong> fetal human neural progenitor<br />

cells (hNPCs). Fetal hNPCs were cultured as three-dimensional neurospheres that proliferate,<br />

migrate and differentiate <strong>in</strong>to both neuronal and glial cells. The migrat<strong>in</strong>g cells from human<br />

neurosphere express functional sGC that synthesize <strong>in</strong>creas<strong>in</strong>g level <strong>of</strong> cGMP upon activation with<br />

NO. Application <strong>of</strong> enzyme <strong>in</strong>hibitors <strong>of</strong> sGC and PKG blocked the migration <strong>of</strong> cells out <strong>of</strong><br />

human neurospheres. Inhibition <strong>of</strong> sGC can be rescued by a membrane permeable analog <strong>of</strong> cGMP.<br />

In ga<strong>in</strong> <strong>of</strong> function experiments both NO donor and cGMP analog facilitate cell migration<br />

suggest<strong>in</strong>g that NO-cGMP signal<strong>in</strong>g positively regulates hNPCs migration. These results provide<br />

first experimental evidence for a role <strong>of</strong> NO/cGMP signal transduction as a regulator <strong>of</strong> cell<br />

migration dur<strong>in</strong>g early development <strong>of</strong> the human bra<strong>in</strong>.<br />

32


Zusammenfassung<br />

Das komplexe Zentralnervensystem (ZNS) der Mammalia durchläuft aufe<strong>in</strong>ander folgende, zum<br />

Teil auch sich überlagernde, zelluläre Entwicklungsprozesse, zu denen Zellproliferation, Migration,<br />

Differenzierung und Synaptogenese gehören. Der gasförmige Botenst<strong>of</strong>f Stickst<strong>of</strong>fmonoxid (NO)<br />

ist mit der Regulation jeder dieser frühen neuronalen Entwicklungsprozesse <strong>in</strong> Zusammenhang<br />

gebracht worden. Allerd<strong>in</strong>gs ist noch nicht bekannt, ob e<strong>in</strong>e NO <strong>Signal</strong>übertragung an der<br />

Entwicklung des humanen ZNS beteiligt ist. In der vorliegenden Arbeit habe ich humane neuronale<br />

Stammzellen als Modell für Nervenzellen <strong>in</strong> der Entwicklung verwendet, um die potentielle Rolle<br />

der NO-<strong>Signal</strong>isierung während der neuronalen Migration, Differenzierung und der präsynaptischen<br />

Vesikel Entleerung zu untersuchen.<br />

Im ersten Teil der vorliegenden Arbeit wurde e<strong>in</strong> Modell für humane neuronale Vorläufer Zellen<br />

aus embryonalen Karz<strong>in</strong>om Stammzellen (NT2) als dreidimensionale, frei flottierende Zellaggregat-<br />

Kultur etabliert. Durch die Behandlung mit Ret<strong>in</strong>säure generieren die sphärischen NT2-Aggregate<br />

Zellen, die proliferieren, migrieren und sich zu neuronalen Phänotypen differenzieren. Dies imitiert<br />

zelluläre Prozesse, die während der frühen Entwicklung des ZNS ablaufen. Interessanterweise<br />

sprechen die migrierenden Zellen und die Zellen <strong>in</strong>nerhalb der Sphären auf exogene Stimulierung<br />

mit NO an, <strong>in</strong>dem sie zyklisches Guanos<strong>in</strong>monophosphat (cGMP) synthetisieren. Dies deutet darauf<br />

h<strong>in</strong>, dass die kugeligen NT2-Aggregate e<strong>in</strong>e funktionale NO/cGMP-<strong>Signal</strong>übertragung aufweisen.<br />

Unter Verwendung von kle<strong>in</strong>en, bioaktiven Enzymaktivatoren und -<strong>in</strong>hibitoren der neuronalen<br />

Stickst<strong>of</strong>fmonoxid Synthase (nNOS), der löslichen Guanylyl Zyklase (sGC) und der Prote<strong>in</strong>k<strong>in</strong>ase<br />

G (PKG) zeigte ich, daß die NO/cGMP/PKG-<strong>Signal</strong>kaskade e<strong>in</strong> positiver Regulator für die<br />

Migration humaner neuronaler Vorläuferzellen ist. Desweiteren konnte e<strong>in</strong> anti-proliferativer Effekt<br />

von NO bei relativ höheren Konzentrationen des NO-Donors beobachtet werden. Außerdem war e<strong>in</strong><br />

leichter, jedoch nicht signifikanter Effekt der NO/cGMP-<strong>Signal</strong>isierung auf die neuronale<br />

Differenzierung zu bemerken.<br />

Im zweiten Teil wurde die präsynaptische Reifung der humanen NT2-Neurone verfolgt, nachdem<br />

diese aus den spärischen NT2-Aggregaten migriert s<strong>in</strong>d und neuronale Netzwerke geformt haben.<br />

<strong>Human</strong>e NT2-Neurone exprimieren ansteigende Spiegel e<strong>in</strong>es axonalen Markers (Tau) und typische<br />

präsynaptische Prote<strong>in</strong>e (Synaps<strong>in</strong> und Synaptotagm<strong>in</strong> I), abhängig von der Dauer der <strong>in</strong> vitro<br />

Kultur. Unter Verwendung e<strong>in</strong>es Antikörpers der gegen die lum<strong>in</strong>ale Domäne von Synaptotagm<strong>in</strong> I<br />

gerichtet ist und e<strong>in</strong>em Fluoreszenzfarbst<strong>of</strong>f (FM1-43) zeigte ich, daß depolarisierte ausgereifte<br />

NT2-Neurone e<strong>in</strong> Calcium-abhängiges, exo-endozytotisches Recycl<strong>in</strong>g der synaptischen Vesikel<br />

aufweisen. NT2-Neurone expremieren nNOS und e<strong>in</strong>en funktionalen NO-Rezeptor, sGC. Als<br />

nächstes untersuchte ich, ob die NO <strong>Signal</strong>transduktion den Turnover der synaptischen Vesikel<br />

33


moduliert. Dieser <strong>Signal</strong>weg könnte zur aktivitätsabhängigen synaptischen Reifung <strong>in</strong> humanen<br />

Modellneuronen beitragen. Zwei NO-Donoren (Natriumnitroprussid und N-Ethyl-2-(1-ethyl-2-<br />

hydroxy-2-nitrosohydraz<strong>in</strong>o ethanam<strong>in</strong>e)) und e<strong>in</strong> cGMP-Analogon (8-Br-cGMP) förderten den<br />

synaptischen Vesikel Turnover, während e<strong>in</strong> sGC-Inhibitor den NO-Donor Effekt blockierte. NO-<br />

Donoren riefen e<strong>in</strong>e Vesikel Exozytose hervor, die teilweise durch den sGC-Inhibitor blockiert<br />

wurde. Forskol<strong>in</strong>, e<strong>in</strong> Aktivator der Adenylyl Zyklase, und e<strong>in</strong> cAMP-Analogon <strong>in</strong>duzierten e<strong>in</strong><br />

Recycl<strong>in</strong>g von synaptischen Vesikeln und Exozytose über e<strong>in</strong>en parallel agierenden Prote<strong>in</strong>k<strong>in</strong>ase<br />

A-<strong>Signal</strong>weg. Diese Ergebnisse deuten an, daß die NO/zyklisches Nukleotid-<strong>Signal</strong>wege die<br />

präsynaptische Neurotransmitter Ausschüttung im humanen Gehirn fördern könnten.<br />

Schließlich habe ich das Vorkommen und die Funktion e<strong>in</strong>er NO/cGMP-<strong>Signal</strong>übertragung <strong>in</strong><br />

fötalen humanen neuronalen Vorläuferzellen (hNPCs) überprüft. Fötale hNPCs wurden als<br />

dreidimensionale Neurosphären kultiviert, welche proliferieren, migrieren und zu neuronalen und<br />

glialen Zellen differenzieren. Die migrierenden Zellen aus humanen Neurosphären exprimieren<br />

funktionale sGCs, welche durch Aktivierung mit NO ansteigende cGMP-Spiegel synthetisieren. Die<br />

Zugabe von Enzym-Inhibitoren von sGC oder PKG blockierten die Zellmigration aus den humanen<br />

Neurosphären. Die Inhibierung der sGC konnte durch e<strong>in</strong> membranpermeables cGMP-Analogon<br />

aufgehoben werden. Sowohl e<strong>in</strong> NO-Donor, als auch e<strong>in</strong> cGMP-Analogon förderte die<br />

Zellmigration <strong>in</strong> ga<strong>in</strong>-<strong>of</strong>-function Experimenten, was nahelegt, daß e<strong>in</strong>e NO-cGMP-<br />

<strong>Signal</strong>übertragung die Migration von hNPCs positiv reguliert. Diese Ergebnisse liefern die ersten<br />

experimentellen Beweise für e<strong>in</strong>e Rolle der NO/cGMP-<strong>Signal</strong>transduktion als e<strong>in</strong> Regulator für die<br />

Zellmigration während der frühen Entwicklung des humanen Gehirns.<br />

34


JOURNAL OF NEUROCHEMISTRY | 2009 | 110 | 1828–1841 doi: 10.1111/j.1471-4159.2009.06279.x<br />

, ,<br />

*Division <strong>of</strong> Cell Biology, Institute <strong>of</strong> Physiology, University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e Hannover, Hannover, Germany<br />

Center for Systems Neuroscience (ZSN), Hannover, Germany<br />

Abstract<br />

Developmental studies <strong>in</strong> both vertebrates and <strong>in</strong>vertebrates<br />

implicate an <strong>in</strong>volvement <strong>of</strong> nitric oxide (NO) signal<strong>in</strong>g <strong>in</strong> cell<br />

proliferation, neuronal motility, and synaptic maturation.<br />

However, it is unknown whether NO plays a role <strong>in</strong> the<br />

development <strong>of</strong> the human nervous system. We used a<br />

model <strong>of</strong> human neuronal precursor cells from a well-characterized<br />

teratocarc<strong>in</strong>oma cell l<strong>in</strong>e (NT2). The precursor cells<br />

proliferate dur<strong>in</strong>g ret<strong>in</strong>oic acid treatment as spherical aggregate<br />

culture that sta<strong>in</strong>s for nest<strong>in</strong> and bIII-tubul<strong>in</strong>. Cells migrate<br />

out <strong>of</strong> the aggregates to acquire fully differentiated<br />

neuronal phenotypes. The cells express neuronal nitric oxide<br />

synthase and soluble guanylyl cyclase (sGC), an enzyme<br />

Migration <strong>of</strong> neuronal progenitors from def<strong>in</strong>ed proliferative<br />

zones to their f<strong>in</strong>al location is a key event that underlies the<br />

development, regeneration, and plasticity <strong>of</strong> the bra<strong>in</strong>.<br />

Dur<strong>in</strong>g early development, newly born neurons undergo<br />

extensive migration to set up the ordered organization <strong>of</strong> the<br />

CNS and PNS. In the adult bra<strong>in</strong>, the subventricular zone<br />

(SVZ) and the subgranular region <strong>of</strong> the hippocampal dentate<br />

gyrus ma<strong>in</strong>ta<strong>in</strong> the capacity <strong>of</strong> generat<strong>in</strong>g new neurons that<br />

migrate out to reach their ultimate locations (Hatten 1999;<br />

Gage 2000; Lambert de Rouvroit and G<strong>of</strong>f<strong>in</strong>et 2001;<br />

Nadarajah et al. 2003; Zheng and Poo 2007). The molecular<br />

mechanisms <strong>of</strong> neuronal migration <strong>in</strong>volve a rearrangement<br />

<strong>of</strong> cytoskeletal components <strong>in</strong> response to extracellular cues,<br />

mediated by numerous <strong>in</strong>tra- and <strong>in</strong>tercellular signals<br />

(reviewed by Ayala et al. 2007; Kawauchi and Hosh<strong>in</strong>o<br />

2008). The gaseous messenger nitric oxide (NO) and its ma<strong>in</strong><br />

target soluble guanylyl cyclase (sGC), an enzyme that<br />

synthesizes cGMP have been implicated <strong>in</strong> neuronal development<br />

<strong>in</strong>clud<strong>in</strong>g neurogenesis and neuronal migrations<br />

(reviewed by Enikolopov et al. 1999; Jurado et al. 2004;<br />

Cárdenas et al. 2005; Bicker 2005, 2007). NO signal<br />

transduction has been shown to differentially regulate cell<br />

proliferation and motility <strong>in</strong> early develop<strong>in</strong>g Xenopus<br />

that synthesizes cGMP upon activation by NO. The migration<br />

<strong>of</strong> the neuronal precursor cell is blocked by the use <strong>of</strong> nNOS,<br />

sGC, and prote<strong>in</strong> k<strong>in</strong>ase G (PKG) <strong>in</strong>hibitors. Inhibition <strong>of</strong><br />

sGC can be rescued by a membrane permeable analog <strong>of</strong><br />

cGMP. In ga<strong>in</strong> <strong>of</strong> function experiments the application <strong>of</strong> a<br />

NO donor and cGMP analog facilitate cell migration. Our<br />

results from the differentiat<strong>in</strong>g NT2 model neurons po<strong>in</strong>t towards<br />

a vital role <strong>of</strong> the NO/cGMP/PKG signal<strong>in</strong>g cascade<br />

as positive regulator <strong>of</strong> cell migration <strong>in</strong> the develop<strong>in</strong>g<br />

human bra<strong>in</strong>.<br />

Keywords: cell migration, cGMP, neural development, nitric<br />

oxide, Ntera 2, prote<strong>in</strong> k<strong>in</strong>ases G.<br />

J. Neurochem. (2009) 110, 1828–1841.<br />

(Peunova et al. 2007). Several studies <strong>in</strong>dicate that NO<br />

enhances neuronal cell migration (Tanaka et al. 1994; Zhang<br />

et al. 2001; Haase and Bicker 2003). The transient expression<br />

<strong>of</strong> NOS dur<strong>in</strong>g development has provided additional<br />

evidence for an <strong>in</strong>volvement <strong>of</strong> NO signal<strong>in</strong>g <strong>in</strong> neuronal<br />

migration. For example, certa<strong>in</strong> neurons <strong>in</strong> the fetal human<br />

sp<strong>in</strong>al cord express a histochemical marker for NOS as they<br />

migrate to their f<strong>in</strong>al dest<strong>in</strong>ation, suggest<strong>in</strong>g a functional role<br />

<strong>of</strong> NO dur<strong>in</strong>g this developmental process (Foster and Phelps<br />

2000). The limited availability <strong>of</strong> fetal tissue will certa<strong>in</strong>ly<br />

Received May 15, 2009; revised manuscript received July 3, 2009;<br />

accepted July 3, 2009.<br />

Address correspondence and repr<strong>in</strong>t requests to Gerd Bicker, Division<br />

<strong>of</strong> Cell Biology, Institute <strong>of</strong> Physiology, University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e<br />

Hannover, Bisch<strong>of</strong>sholer Damm 15, D-30173 Hannover, Germany.<br />

E-mail: gerd.bicker@tiho-hannover.de<br />

Abbrevations used: 7NI, 7-nitro<strong>in</strong>dazole; 8-Br-cGMP, 8-bromo-cylic<br />

guanos<strong>in</strong>e-monophosphate; BrdU, 5-bromo-2¢-deoxyurid<strong>in</strong>e; cGMP-IR,<br />

cGMP immunoreactivity; DMEM, Dulbecco’s modified eagle medium;<br />

nNOS, neuronal nitric oxide synthase; NO, nitric oxide; ODQ, 1H-<br />

[1,2,4]-oxadiazolo[4,3-a]qu<strong>in</strong>oxal<strong>in</strong>-1-one; PBS, phosphate-buffered<br />

sal<strong>in</strong>e; PDL, poly-D-lys<strong>in</strong>e; PKG, prote<strong>in</strong> k<strong>in</strong>ase G; PTX, PBS conta<strong>in</strong><strong>in</strong>g<br />

0.2% Triton X-100; RA, ret<strong>in</strong>oic acid; sGC, soluble guanylyl<br />

cyclase; SNP, sodium nitroprusside; SVZ, subventricular zone.<br />

Ó 2009 The Authors<br />

1828 Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841


estrict experimental approaches to unravel the function <strong>of</strong><br />

NO signal<strong>in</strong>g <strong>in</strong> human bra<strong>in</strong> development.<br />

The human teratocarc<strong>in</strong>oma cell l<strong>in</strong>e Ntera2 (NT2) is<br />

derived from testicular cancer that can be <strong>in</strong>duced to<br />

differentiate <strong>in</strong>to fully functional, post-mitotic neurons<br />

(Andrews 1984; Pleasure et al. 1992). Evidence from gene<br />

pr<strong>of</strong>il<strong>in</strong>g <strong>of</strong> NT2 cells treated with ret<strong>in</strong>oic acid (RA)<br />

<strong>in</strong>dicates the similarity <strong>of</strong> NT2 neural differentiation to<br />

vertebrate neurogenesis (Przyborski et al. 2000, 2003; Houldsworth<br />

et al. 2002). Moreover, NT2 neurons have been<br />

successfully transplanted <strong>in</strong> human stroke patients (Nelson<br />

et al. 2002; Hara et al. 2008). Recent studies from our group<br />

have shown that NT2 cells could be <strong>in</strong>duced to differentiate<br />

<strong>in</strong>to post-mitotic neurons upon RA treatment <strong>in</strong> spherical<br />

aggregate culture and express a number <strong>of</strong> neuronal markers<br />

(Paquet-Durand et al. 2003; Podrygajlo et al. 2009).<br />

In this study we used NT2 spherical aggregates as a model<br />

<strong>of</strong> human neuronal precursor cells to elucidate the <strong>in</strong>volvement<br />

<strong>of</strong> NO-cGMP signal<strong>in</strong>g <strong>in</strong> neuronal precursor cell<br />

migration. The NT2 spherical aggregates were characterized<br />

by immunocytochemical methods for the expression <strong>of</strong><br />

neuronal markers, the neuronal is<strong>of</strong>orm <strong>of</strong> the NO synthesiz<strong>in</strong>g<br />

enzyme, neuronal nitric oxide synthase (nNOS), and<br />

its target enzyme, sGC. Application <strong>of</strong> enzyme <strong>in</strong>hibitors and<br />

activators to the differentiat<strong>in</strong>g NT2 cell aggregate provides<br />

firm evidence that the NO/cGMP/ prote<strong>in</strong> k<strong>in</strong>ase G (PKG)<br />

signal transduction pathway positively regulates human<br />

neuronal precursor cell migration.<br />

Materials and methods<br />

Materials<br />

The NO donor, NOC-18 (2,2-(Hydroxynitrosohydraz<strong>in</strong>o) bis-ethanam<strong>in</strong>e]),<br />

was purchased from Calbiochem (Darmstadt, Germany),<br />

and 8-bromo-cylic guanos<strong>in</strong>e-monophosphate (8-Br-cGMP) and RP<br />

isomer <strong>of</strong> 8-Br-cGMP were purchased from Alexis Biochemicals<br />

(Lörrach, Germany). All the antibodies and other substances were<br />

obta<strong>in</strong>ed from Sigma (Taufkirchen, Germany) unless otherwise<br />

noted.<br />

Cell culture<br />

NT2/D1 precursor cells were purchased from American Type Culture<br />

Collection (ATTC, Manassas, VA, USA) and treated as <strong>in</strong>dicated <strong>in</strong><br />

the ATTC <strong>in</strong>struction manual. NT2 cells were cultured as free float<strong>in</strong>g<br />

spherical aggregate as described by Paquet-Durand et al. (2003).<br />

Briefly, the NT2 precursor cells (passages 24–32) were seeded <strong>in</strong><br />

80 mm, bacteriological grade Petri dishes (Gre<strong>in</strong>er, Hamburg, FRG)<br />

at a density <strong>of</strong> 5 · 10 6 cells/dish <strong>in</strong> 10 mL <strong>of</strong> Dulbecco’s modified<br />

eagle medium (DMEM/F-12; Gibco–Invitrogen, Karlsruhe, Germany)<br />

supplemented with 10% fetal bov<strong>in</strong>e serum (Invitrogen,<br />

Karlsruhe, Germany) for 24 h. A differentiation medium, DMEM<br />

conta<strong>in</strong><strong>in</strong>g RA at a f<strong>in</strong>al concentration <strong>of</strong> 10 lM was used with<br />

medium change every 2–3 days. One group <strong>of</strong> spherical aggregates<br />

were cultured <strong>in</strong> a dish with RA-conta<strong>in</strong><strong>in</strong>g medium for 7 days<br />

(1 week) and the other group for 14 days (2 weeks).<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841<br />

NO signal<strong>in</strong>g regulates neuronal precursor motility | 1829<br />

Migration assay<br />

The spherical aggregate culture generated <strong>in</strong> a Petri dish was seeded<br />

<strong>in</strong>to poly-D-lys<strong>in</strong>e (PDL) and Matrigel (Becton-Dick<strong>in</strong>son, Bedford,<br />

MA, USA) coated microdishes (Ibidi GmbH, Munich, Germany) or<br />

12-mm cover glasses at a density <strong>of</strong> approximately 2–10 aggregates.<br />

The culture was allowed to attach for 90 m<strong>in</strong> <strong>in</strong> an <strong>in</strong>cubator at<br />

37°C/5% CO 2. NOC-18 was prepared <strong>in</strong> 10 mM NaOH as 100 mM<br />

stock solution, 1H-[1,2,4]-oxadiazolo[4,3-a]qu<strong>in</strong>oxal<strong>in</strong>-1-one (ODQ)<br />

dissolved <strong>in</strong> dimethylsulfoxide as 20 mM stock, 7-nitro<strong>in</strong>dazole (7NI)<br />

dissolved <strong>in</strong> dimethylsulfoxide as 200 mM stock, and 8-Br-cGMP<br />

and RP-8-Br-cGMP were dissolved <strong>in</strong> the medium. DMEM-conta<strong>in</strong><strong>in</strong>g<br />

RA and chemical compounds were added <strong>in</strong>to the culture and<br />

<strong>in</strong>cubated at 37°C/5% CO 2 for 48 h. To determ<strong>in</strong>e the migration <strong>of</strong><br />

cells out <strong>of</strong> the spherical aggregate we acquired images at desired<br />

times as described <strong>in</strong> Moors et al. (2007).<br />

Immunocytochemistry<br />

For immunocytochemical sta<strong>in</strong><strong>in</strong>gs, we either used spherical aggregate<br />

cultures after 48 h <strong>of</strong> the migration assay or aggregates that were<br />

mechanically dispersed <strong>in</strong>to s<strong>in</strong>gle cells. The cultures were fixed <strong>in</strong><br />

4% paraformaldehyde dissolved <strong>in</strong> phosphate-buffered sal<strong>in</strong>e (PBS,<br />

10 mM sodium phosphate, 150 mM NaCl, pH 7.4) for 30 m<strong>in</strong>. They<br />

were permeabilized by wash<strong>in</strong>g three times for 5 m<strong>in</strong> <strong>in</strong> PBS<br />

conta<strong>in</strong><strong>in</strong>g 0.2% Triton X-100 (PTX). Block<strong>in</strong>g solution conta<strong>in</strong><strong>in</strong>g<br />

PTX and 5% normal horse or rabbit serum was applied for 1 h. The<br />

primary anti-bIII-tubul<strong>in</strong> (1 : 10000; Sigma) and anti-nest<strong>in</strong><br />

(1 : 400; Calbiochem) were diluted <strong>in</strong> PTX-conta<strong>in</strong><strong>in</strong>g block<strong>in</strong>g<br />

solution and applied overnight at 4°C. Wells were washed three times<br />

for 5 m<strong>in</strong> <strong>in</strong> PTX and <strong>in</strong>cubated with secondary biot<strong>in</strong>ylated antibody<br />

(Vector, Burl<strong>in</strong>game, CA, USA) for 1 h at 20°C. After three more<br />

wash<strong>in</strong>g steps <strong>in</strong> PTX, the immun<strong>of</strong>luorescence was detected us<strong>in</strong>g<br />

streptavid<strong>in</strong>-CY3 (Sigma) or streptavid<strong>in</strong>-Alexa Fluor 488 (Mobitec,<br />

Gött<strong>in</strong>gen, Germany). For nuclear counter-sta<strong>in</strong><strong>in</strong>g cells were<br />

<strong>in</strong>cubated with 4¢-6-diamid<strong>in</strong>o-2-phenyl<strong>in</strong>dole at a concentration <strong>of</strong><br />

1 lg/mL for 5 m<strong>in</strong> <strong>in</strong> PBS. After f<strong>in</strong>al wash<strong>in</strong>g steps <strong>in</strong> PBS,<br />

preparations were mounted <strong>in</strong> 90/10% glycerol/PBS with 4% sodium<br />

n-propyl-gallate as antifad<strong>in</strong>g agent and sealed with nail varnish.<br />

For the detection <strong>of</strong> cGMP immunoreactivity (cGMP-IR),<br />

spherical aggregates were pre-<strong>in</strong>cubated for 20 m<strong>in</strong> at 20°C with<br />

1 mM sodium nitroprusside (SNP) as a NO donor, 20 lM YC-1 [3-<br />

(50-Hydroxymethyl-20-furyl)-1-benzyl <strong>in</strong>dazole] as an enhancer <strong>of</strong><br />

NO-<strong>in</strong>duced activity <strong>of</strong> sGC, and 1 mM 3-isobutyl-1-methylxanth<strong>in</strong>e<br />

as phosphodiesterase <strong>in</strong>hibitor. Cultures were washed once<br />

with PBS and then we followed the same sta<strong>in</strong><strong>in</strong>g procedures as<br />

above with a block<strong>in</strong>g solution conta<strong>in</strong><strong>in</strong>g PTX and 5% normal<br />

rabbit serum. The polyclonal sheep cGMP antiserum (1 : 10000; a<br />

k<strong>in</strong>d gift from Dr. J. de Vente, Maastricht University, Netherlands)<br />

was used as primary antibody to detect the level <strong>of</strong> cGMP. To test<br />

for the contribution <strong>of</strong> endogenous enzyme activity to cGMP levels,<br />

we used 50 lM ODQ as a specific <strong>in</strong>hibitor <strong>of</strong> sGC.<br />

Western blott<strong>in</strong>g<br />

Cells were homogenized <strong>in</strong> a ristocet<strong>in</strong>-<strong>in</strong>duced platelet agglut<strong>in</strong>ation<br />

lysis buffer and HALT Ò protease <strong>in</strong>hibitor cocktail (Pierce,<br />

Rockford, IL, USA) for about 30 m<strong>in</strong>. The cell lysates were<br />

centrifuged at speed <strong>of</strong> 6000 g for 10 m<strong>in</strong> and the prote<strong>in</strong> level <strong>of</strong><br />

supernatant was estimated by the BCA TM prote<strong>in</strong> assay kit (Pierce).<br />

Subsequently, equal amount <strong>of</strong> prote<strong>in</strong> (100 lg) for NT2 and


1830 | M. A. Tegenge and G. Bicker<br />

NT2 + RA-treated samples were boiled for 3 m<strong>in</strong> <strong>in</strong> 2· Laemmli<br />

buffer and subjected to electrophoresis on 8% sodium dodecyl sulfate<br />

acrylamide gel. The gel was blotted on polyv<strong>in</strong>ylidene difloride<br />

membrane at 4°C. The membrane was then blocked with 3% dried<br />

milk powder <strong>in</strong> PBS for 2 h and <strong>in</strong>cubated overnight at 4°C with a<br />

monoclonal antibody aga<strong>in</strong>st nNOS (Sigma) at 1 : 1000 dilution <strong>in</strong><br />

block<strong>in</strong>g solution. The biot<strong>in</strong>ylated secondary antibody was applied<br />

for 1 h. Bound antiserum was detected by a standard peroxidase<br />

sta<strong>in</strong><strong>in</strong>g techniques us<strong>in</strong>g the Vectasta<strong>in</strong> ABC Kit (Vector). After<br />

reactivation with methanol, membranes were additionally sta<strong>in</strong>ed for<br />

anti-acetylated-a-tubul<strong>in</strong> diluted 1 : 10000 <strong>in</strong> block<strong>in</strong>g solution.<br />

Proliferation assay<br />

Cell proliferation was assessed by application <strong>of</strong> the thymid<strong>in</strong>e<br />

analog 5-bromo-2¢-deoxyurid<strong>in</strong>e (BrdU), which was <strong>in</strong>corporated<br />

<strong>in</strong>to the DNA <strong>of</strong> divid<strong>in</strong>g cells. Spherical aggregates were exposed<br />

to chemicals for 48 h and then <strong>in</strong>cubated with BrdU (50 lM) for<br />

4 h. After fixation with 4% paraformaldehyde for 30 m<strong>in</strong>, they were<br />

<strong>in</strong>cubated with 2 M HCl solution <strong>in</strong> PBS for 20 m<strong>in</strong>. Preparations<br />

were washed three times with PBS to completely remove the acid.<br />

They were permeabilized by wash<strong>in</strong>g three times for 5 m<strong>in</strong> <strong>in</strong> 0.2%<br />

PTX and blocked for 1 h <strong>in</strong> 5% normal horse serum. Then, we<br />

followed the same immunocytochemical sta<strong>in</strong><strong>in</strong>g procedures as<br />

described <strong>in</strong> the Immunocytochemistry with the primary monoclonal<br />

anti-BrdU (clone Bu-33, Sigma) 1 : 200 <strong>in</strong> block<strong>in</strong>g solution.<br />

Cell viability/cytotoxicity assay<br />

To monitor cytotoxic effects <strong>of</strong> the drugs, we used the Alamar Blue<br />

viability assay (Trek Diagnostic Systems, East Gr<strong>in</strong>stead, UK) and<br />

the Live/Dead viability/cytotoxicity assay (Molecular Probes,<br />

Eugene, OR, USA). For Alamar Blue viability assay the NT2<br />

spherical aggregates were mechanically dispersed <strong>in</strong>to s<strong>in</strong>gle cells.<br />

About 10000 cell/well were seeded <strong>in</strong>to PDL and Matrigel-coated<br />

96-well plate and allowed to attach for 90 m<strong>in</strong>. RA and chemical<br />

compounds conta<strong>in</strong><strong>in</strong>g medium were added and <strong>in</strong>cubated for 48 h at<br />

37°C/5% CO 2. After 48 h the medium was completely changed <strong>in</strong>to<br />

a cell culture medium conta<strong>in</strong><strong>in</strong>g 3% Alamar Blue. Fluorescence<br />

<strong>in</strong>tensity <strong>of</strong> the Alamar Blue was detected at excitation/emission<br />

wavelength <strong>of</strong> 530/590 nm us<strong>in</strong>g a microplate reader (Inf<strong>in</strong>ite M200,<br />

Tecan, Männedorf, Switzerland) after 4 h. The actual number <strong>of</strong><br />

liv<strong>in</strong>g or dead cells was determ<strong>in</strong>ed by <strong>in</strong>cubat<strong>in</strong>g NT2 aggregates<br />

with 4 lM <strong>of</strong> ethidium homodimer-1 (EthD-1) and 2 lM calce<strong>in</strong> <strong>in</strong><br />

PBS for 30 m<strong>in</strong> at 37°C/5% CO 2. After <strong>in</strong>cubation the cultures were<br />

washed <strong>in</strong> PBS and viewed under microscope. The average numbers<br />

<strong>of</strong> liv<strong>in</strong>g and dead cells were calculated from at least 10 images <strong>of</strong><br />

two <strong>in</strong>dependent experiments and compared with live controls which<br />

was normalized to 100% for each experiment.<br />

Microscopy and data analysis<br />

The morphological observation and migration assay were performed<br />

on a Zeiss Axiovert 200 microscope (Zeiss, Gött<strong>in</strong>gen, Germany)<br />

equipped with a CoolSNAP digital camera. Immun<strong>of</strong>luorescence<br />

was detected on a Zeiss Axiovert 25 microscope equipped with an<br />

Axiocam 3900 digital camera. The acquired images were processed<br />

us<strong>in</strong>g Adobe Photoshop (Adobe Systems GmbH, München,<br />

Germany). The migration was quantified by measur<strong>in</strong>g the distance<br />

from the edge <strong>of</strong> the sphere to the furthest migrated cell bodies at<br />

four different locations per spherical aggregate (Moors et al. 2007).<br />

The data were presented as mean ± SEM <strong>of</strong> at least eight spherical<br />

aggregate. Each experiment was repeated at least three times.<br />

The percentage <strong>of</strong> nest<strong>in</strong>, bIII-tubul<strong>in</strong>, and BrdU were determ<strong>in</strong>ed<br />

divid<strong>in</strong>g the number <strong>of</strong> positively sta<strong>in</strong>ed cells by the total<br />

number <strong>of</strong> cells. The data were presented as mean ± SEM <strong>of</strong> two<br />

<strong>in</strong>dependent experiments. Statistical comparisons <strong>of</strong> controls versus<br />

treatment were performed with the unpaired two-tailed Student’s<br />

t-test. Levels <strong>of</strong> significance were <strong>in</strong>dicated as *p < 0.05,<br />

**p < 0.01, and ***p < 0.001.<br />

Results<br />

Migration <strong>of</strong> cells out <strong>of</strong> the NT2 spherical aggregates<br />

depends on the length <strong>of</strong> RA treatment<br />

After form<strong>in</strong>g confluent monolayers NT2 precursor cells<br />

were split and seeded <strong>in</strong>to bacteriological dishes. The cells<br />

formed a loose aggregate after 24 h <strong>of</strong> the proliferation step<br />

followed by subsequent RA treatment for 2 days. Upon<br />

further <strong>in</strong>cubation with medium conta<strong>in</strong><strong>in</strong>g RA for 1 week<br />

the aggregates became spherical <strong>in</strong> shape, stable, and<br />

surrounded by fewer isolated cells. When the use <strong>of</strong> RA<br />

was prolonged for 2 weeks, the aggregates atta<strong>in</strong>ed a more<br />

spherical shape and elaborated neuronal processes. Accord<strong>in</strong>gly,<br />

we classified the aggregates <strong>in</strong>to two separate groups<br />

which were based on age as 1 or 2 weeks cell spheres.<br />

The aggregates were seeded <strong>in</strong>to PDL and Matrigel-coated<br />

microdishes or cover glasses conta<strong>in</strong><strong>in</strong>g a marker grid, which<br />

allowed us to follow the migration <strong>of</strong> cells out <strong>of</strong> the same<br />

sphere for 48 h. Cells started to detach radially from the rim <strong>of</strong><br />

1-week-old NT2 sphere after about 3–6 h and the migrated<br />

distance <strong>in</strong>creased with time (Fig. 1a–c). Cells migrated to a<br />

lesser extent from 2-week-old NT2 spheres but elaborated<br />

neuronal processes (Fig. 1d–f). Higher magnification <strong>of</strong> the<br />

migrated area <strong>of</strong> 1-week-old spheres showed mostly s<strong>in</strong>gle<br />

cells devoid <strong>of</strong> neuronal processes (Fig. 1g). However, 2week-old<br />

spheres elaborated long neuronal processes (Fig. 1h)<br />

and <strong>in</strong> this case length <strong>of</strong> neurites was excluded from the<br />

migration analysis. Accord<strong>in</strong>gly, migratory cells derived from<br />

1- and 2-week-old spheres cover, after 48 h, a distance <strong>of</strong><br />

171.0 ± 5.5 and 77.5 ± 2.1 lm, respectively (Fig. 1i). Thus,<br />

the migration <strong>of</strong> cells out <strong>of</strong> the NT2 spheres depended on the<br />

duration <strong>of</strong> the RA treatment <strong>in</strong> the Petri dishes.<br />

NT2 spheres express both precursor and neuronal markers<br />

To monitor the progress <strong>of</strong> neuronal differentiation with<strong>in</strong><br />

the develop<strong>in</strong>g spherical aggregate, we performed immunocytochemical<br />

sta<strong>in</strong><strong>in</strong>g aga<strong>in</strong>st cytoskeletal markers <strong>of</strong> progenitor<br />

cells and neurons. At the end <strong>of</strong> a 48 h last<strong>in</strong>g<br />

migration experiment spheres were fixed and sta<strong>in</strong>ed aga<strong>in</strong>st<br />

the neuronal progenitor marker nest<strong>in</strong> and an early marker <strong>of</strong><br />

neuronal differentiation, bIII-tubul<strong>in</strong>. After 1 week <strong>of</strong> RA<br />

treatment the NT2 spheres expressed nest<strong>in</strong> and bIII-tubul<strong>in</strong><br />

(Fig. 2a and c). For quantification, the spheres were<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841


Fig. 1 Cells migrate out <strong>of</strong> NT2 spheres<br />

that were seeded on PDL and Matrigelcoated<br />

microdishes. Representative photomicrographs<br />

<strong>of</strong> NT2 spheres treated with<br />

10 lM RA for 1 week (a–c) and 2 week (d–<br />

f). The white l<strong>in</strong>es <strong>in</strong> c del<strong>in</strong>eate distance <strong>of</strong><br />

migration from the rim <strong>of</strong> the sphere to the<br />

furthest migrated cells. Arrow heads <strong>in</strong> (f)<br />

show representative neuronal processes. A<br />

higher magnification <strong>of</strong> the migrated cells<br />

shows s<strong>in</strong>gle cells that bear no or few<br />

neurite <strong>in</strong> 1-week-old spheres (g) when<br />

compared with 2-week-old spheres that<br />

elaborated long neurites (h). The use <strong>of</strong> RA<br />

for 2 weeks significantly decreased cell<br />

migration as compared with 1-week-old<br />

spheres (i). Scale bars: 100 lm<br />

(***p < 0.001).<br />

dispersed mechanically <strong>in</strong>to s<strong>in</strong>gle cells and sta<strong>in</strong>ed<br />

aga<strong>in</strong>st both markers. While 10.3% <strong>of</strong> the cells sta<strong>in</strong>ed<br />

positively for bIII-tubul<strong>in</strong> (Fig. 2e and h), about 84.3%<br />

rema<strong>in</strong>ed nest<strong>in</strong>-positive (Fig. 2g). Upon further <strong>in</strong>cubation<br />

with RA for an additional week, the spheres displayed<br />

an <strong>in</strong>creas<strong>in</strong>g amount <strong>of</strong> bIII-tubul<strong>in</strong>-sta<strong>in</strong>ed cells (Fig. 2d).<br />

The immunoreactivity was not only expressed <strong>in</strong> the cell<br />

bodies but also <strong>in</strong> the emerg<strong>in</strong>g neuronal processes (Fig. 2d<br />

and f). We counted about 46% bIII-tubul<strong>in</strong>- and 63%<br />

nest<strong>in</strong>-positive cells (Fig. 2g and h). An <strong>in</strong>crease <strong>in</strong> the<br />

level <strong>of</strong> bIII-tubul<strong>in</strong>, elaboration <strong>of</strong> neuronal processes<br />

together with reduction <strong>in</strong> the level <strong>of</strong> nest<strong>in</strong> upon longer<br />

<strong>in</strong>cubation with RA <strong>in</strong>dicated the progress <strong>of</strong> neuronal<br />

differentiation.<br />

(a) (b) (c)<br />

(d) (e)<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841<br />

(g)<br />

(i)<br />

NO signal<strong>in</strong>g regulates neuronal precursor motility | 1831<br />

(h)<br />

Studies on gene expression <strong>of</strong> NT2 cell dur<strong>in</strong>g and<br />

shortly after differentiation with RA showed down-regulations<br />

<strong>of</strong> several stem cell markers and up-regulation <strong>of</strong><br />

nest<strong>in</strong> and neur<strong>of</strong>ilament transcripts that support our present<br />

observations at the prote<strong>in</strong> levels (Houldsworth et al. 2002;<br />

Przyborski et al. 2003). Because <strong>of</strong> their high cellular<br />

motility we further characterized the type <strong>of</strong> cells that<br />

migrated out <strong>of</strong> 1-week-old NT2 spheres. Only 3.5 ± 0.5%<br />

<strong>of</strong> the migratory cells were bIII-tubul<strong>in</strong>-positive and almost<br />

all <strong>of</strong> the furthest migrated cells lacked the bIII-tubul<strong>in</strong><br />

sta<strong>in</strong><strong>in</strong>g (Fig. 2c). Hence, the cells at the front <strong>of</strong> the<br />

migration expressed ma<strong>in</strong>ly the precursor marker nest<strong>in</strong>.<br />

Costa<strong>in</strong><strong>in</strong>g <strong>of</strong> BrdU together with nest<strong>in</strong> <strong>in</strong>dicated that most<br />

<strong>of</strong> the proliferat<strong>in</strong>g cells were nest<strong>in</strong>-positive neuronal<br />

(f)


1832 | M. A. Tegenge and G. Bicker<br />

(a) (b)<br />

(c) (d)<br />

(e) (f)<br />

(g) (h)<br />

progenitor cells (Fig. S1a). Costa<strong>in</strong><strong>in</strong>g <strong>of</strong> BrdU together<br />

with bIII-tubul<strong>in</strong> showed that none <strong>of</strong> the BrdU-<strong>in</strong>corporated<br />

cells were bIII-tubul<strong>in</strong> positive (Fig. S1b). Thus,<br />

Fig. 2 NT2 cell spheres express both nest<strong>in</strong><br />

and bIII-tubul<strong>in</strong>. NT2 cell spheres were<br />

treated with 10 lM RA for a period <strong>of</strong> 1<br />

and 2 weeks with medium change every<br />

2–3 days. At the end <strong>of</strong> a 48 h migration<br />

experiment, spheres were fixed and sta<strong>in</strong>ed<br />

aga<strong>in</strong>st nest<strong>in</strong> and bIII-tubul<strong>in</strong>. One-weekold<br />

NT2 spheres (a) express higher level <strong>of</strong><br />

nest<strong>in</strong> (green) than 2-week-old spheres (b).<br />

The expression <strong>of</strong> bIII-tubul<strong>in</strong> <strong>in</strong> cell soma<br />

and neuronal processes (red) after 2 weeks<br />

RA treatment (d) <strong>in</strong>creases <strong>in</strong> comparison<br />

with the 1-week-old counterparts (c). Panels<br />

(e–h) were obta<strong>in</strong>ed from spheres that<br />

were mechanically dispersed <strong>in</strong>to s<strong>in</strong>gle<br />

cells. Quantification <strong>of</strong> positively sta<strong>in</strong>ed<br />

cells showed that upon RA treatment for<br />

2 weeks the level <strong>of</strong> nest<strong>in</strong> (g) was reduced<br />

while bIII-tubul<strong>in</strong> (h) was significantly<br />

<strong>in</strong>creased. Blue (4¢-6-diamid<strong>in</strong>o-2-phenyl<strong>in</strong>dole)<br />

<strong>in</strong>dicates nuclear counter-sta<strong>in</strong><strong>in</strong>g.<br />

Scale bars: 100 lm (*p < 0.05, ***p <<br />

0.001).<br />

migration as measured <strong>in</strong> our cell culture model was ma<strong>in</strong>ly<br />

based on the motility <strong>of</strong> nest<strong>in</strong>-positive neural precursor<br />

cells (Fig. 2a). Along these l<strong>in</strong>es we used the NT2 cell<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841


spheres to study the role <strong>of</strong> NO signal<strong>in</strong>g <strong>in</strong> neuronal<br />

precursor migration.<br />

Agents that modulate cell migration via NO signal<strong>in</strong>g do<br />

not alter cell viability and differentiation<br />

We <strong>in</strong>itially exam<strong>in</strong>ed the potential cytotoxicity <strong>of</strong> bioactive<br />

chemical compounds affect<strong>in</strong>g cell migration via NO/cGMP<br />

signal transduction (Fig. 3a). Thus, we performed Alamar<br />

Blue cell viability and the live/dead assay under identical<br />

Fig. 3 Schematic draw<strong>in</strong>g <strong>of</strong> NO signal<br />

transduction together with an array <strong>of</strong><br />

chemical compounds that block or activate<br />

the pathway (a). The enzyme nitric oxide<br />

synthase (nNOS) is stimulated by calcium–<br />

calmodul<strong>in</strong> (Ca/CaM) and can be blocked<br />

by bath application <strong>of</strong> the <strong>in</strong>hibitor, 7-nitro<strong>in</strong>dazole<br />

(7NI). NO b<strong>in</strong>ds to the heme moiety<br />

<strong>of</strong> soluble guanylyl cyclase (sGC)<br />

result<strong>in</strong>g <strong>in</strong> the stimulation <strong>of</strong> the enzyme.<br />

The sGC activity is blocked by the <strong>in</strong>hibitor<br />

ODQ and stimulated <strong>in</strong>dependently from<br />

NO by the sGC activator YC-1. Synthesis <strong>of</strong><br />

cGMP activates prote<strong>in</strong> k<strong>in</strong>ase G (PKG)<br />

and regulates cell migration. The PKG<br />

<strong>in</strong>hibitor, RP-8-Br-cGMP, blocks cellular<br />

responses <strong>of</strong> the cGMP/PKG pathway. The<br />

NO donors (SNP and NOC-18) and the<br />

membrane-permeable cGMP analog (8-BrcGMP)<br />

can be applied to raise cGMP<br />

levels. The Live/Dead and Alamar Blue<br />

viability/cytotoxicity assay was performed <strong>in</strong><br />

the presence <strong>of</strong> chemical compounds that<br />

modulate the NO/cGMP signal<strong>in</strong>g. Application<br />

<strong>of</strong> 7NI (500 lM), ODQ (50 lM), RP-8-<br />

Br-cGMP (100 nM), NOC-18 (50 lM), and<br />

8-Br-cGMP (1 mM) to NT2 spheres did not<br />

affect cell viability as determ<strong>in</strong>ed by Live/<br />

Dead assay (b) and Alamar Blue assay (c).<br />

Relatively higher concentrations (1 mM) <strong>of</strong><br />

7NI, NOC-18, and ODQ significantly affected<br />

cell viability as determ<strong>in</strong>ed by Alamar<br />

Blue assay (d–f). Application <strong>of</strong> 7NI<br />

(500 lM), ODQ (50 lM), RP-8-Br-cGMP<br />

(100 nM), NOC-18 (50 lM), and 8-BrcGMP<br />

(1 mM) to NT2 spheres did not significantly<br />

alter the bIII-tubul<strong>in</strong>-sta<strong>in</strong>ed cells<br />

(g). NOC-18, 2, 2-hydroxynitrosohydraz<strong>in</strong>obis-ethanam<strong>in</strong>e;<br />

RP-8-Br-cGMP, RP isomer<br />

<strong>of</strong> 8-Br-cGMP; Ctrl, control (***p < 0.001).<br />

(a)<br />

experimental conditions as the migration assay. Both assay<br />

showed that none <strong>of</strong> the pharmacological agents had any<br />

effect on cell viability at the concentration used to modulate<br />

cell migration (Fig. 3b and c). This was also further<br />

confirmed by morphological observation <strong>of</strong> the spheres<br />

<strong>in</strong>clud<strong>in</strong>g the migrat<strong>in</strong>g cells <strong>in</strong> the presence <strong>of</strong> the<br />

compounds. However, when the NOS <strong>in</strong>hibitor, 7NI; sGC<br />

<strong>in</strong>hibitor; ODQ; and NO donor, NOC-18 were used at the<br />

rather high concentration <strong>of</strong> 1 mM, they significantly<br />

(b) (c)<br />

(d) (e)<br />

(f) (g)<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841<br />

NO signal<strong>in</strong>g regulates neuronal precursor motility | 1833


1834 | M. A. Tegenge and G. Bicker<br />

affected the survival <strong>of</strong> cells (Fig. 3d–f). Next, we used the<br />

bIII-tubul<strong>in</strong> immunocytochemical sta<strong>in</strong><strong>in</strong>g to determ<strong>in</strong>e<br />

the effect <strong>of</strong> chemicals on neuronal differentiation. At the<br />

concentration used to modulate cell migration, only 7NI<br />

appeared to slightly decrease while 8-Br-cGMP <strong>in</strong>creased the<br />

bIII-tubul<strong>in</strong> sta<strong>in</strong><strong>in</strong>g. However, the effect <strong>of</strong> both chemicals<br />

was not statistically significant (Fig. 3g).<br />

NT2 cells express nNOS and functional sGC<br />

Western blott<strong>in</strong>g revealed the presence <strong>of</strong> the nNOS <strong>in</strong> both<br />

RA-treated and untreated precursor cells (Fig. 4a). Us<strong>in</strong>g a<br />

(a)<br />

NT2 + RA<br />

NT2<br />

155 kD<br />

55 kD<br />

(b) Ctrl (c)<br />

(d) SNP + ODQ (e)<br />

ODQ<br />

monoclonal antibody aga<strong>in</strong>st nNOS, the blot resolved a<br />

major prote<strong>in</strong> band at 155 kDa. This result is <strong>in</strong> l<strong>in</strong>e with<br />

studies show<strong>in</strong>g that untreated NT2 cells express nNOS but<br />

not endothelial NOS and <strong>in</strong>ducible NOS (Lee et al. 2001;<br />

Hyun et al. 2002). For comparison <strong>of</strong> the NOS content <strong>in</strong><br />

precursor NT2 with NT2 + RA-treated cells, the ratio <strong>of</strong><br />

nNOS signal to the load<strong>in</strong>g control was calculated and<br />

averaged for all probed blots (n = 3). We found no<br />

significant differences between NT2 + RA and NT2 cells<br />

(0.76 ± 0.07 vs. 0.68 ± 0.11, respectively). To reveal<br />

potential cellular targets <strong>of</strong> a NO signal endogenous to the<br />

SNP<br />

Fig. 4 NT2 cells express nNOS and functional<br />

sGC. Western blott<strong>in</strong>g <strong>of</strong> lysates from<br />

NT2 cell spheres treated with RA for<br />

1 week and untreated NT2 cells. The antibody<br />

aga<strong>in</strong>st nNOS recognizes a prote<strong>in</strong><br />

band <strong>of</strong> apparent molecular weight at<br />

155 kDa (a). Lower band represents a<br />

load<strong>in</strong>g control probed with acetylated-atubul<strong>in</strong><br />

antibody. Immuncytochemical<br />

sta<strong>in</strong><strong>in</strong>g <strong>of</strong> cell spheres for cGMP levels (b–<br />

e). After 48 h <strong>of</strong> seed<strong>in</strong>g the spheres <strong>in</strong><br />

PDL- and Matrigel-coated cover glasses,<br />

the cultures were exposed for 20 m<strong>in</strong> to<br />

SNP (1 mM), ODQ (50 lM), or a comb<strong>in</strong>ation<br />

<strong>of</strong> SNP and ODQ together with 1 mM<br />

3-isobutyl-1-methylxanth<strong>in</strong>e (IBMX) and<br />

20 lM YC-1. Controls were also treated<br />

with IBMX and YC-1. Cultures were fixed<br />

with 4% PFA and sta<strong>in</strong>ed aga<strong>in</strong>st cGMP.<br />

NT2 cell sphere shows very low cGMP-IR<br />

(b). The addition <strong>of</strong> NO donor SNP <strong>in</strong>creases<br />

the cGMP-IR (c). The use <strong>of</strong> SNP<br />

together with ODQ resulted <strong>in</strong> reduced<br />

cGMP-IR (d). The cGMP-IR is almost<br />

abolished <strong>in</strong> the presence <strong>of</strong> sGC <strong>in</strong>hibitor,<br />

ODQ (e). Scale bar: 100 lm. Ctrl, control.<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841


NT2 cell spheres, we used an antibody aga<strong>in</strong>st cGMP (de<br />

Vente et al. 1987; Tanaka et al. 1997). In the absence <strong>of</strong> an<br />

exogenous source <strong>of</strong> NO, a low level <strong>of</strong> cGMP was detected<br />

which further confirmed the presence <strong>of</strong> an endogenous<br />

source <strong>of</strong> NO (Fig. 4b). The level <strong>of</strong> cGMP <strong>in</strong>creased<br />

dramatically upon stimulation with the NO donor SNP<br />

(Fig. 4c). We could not observe the <strong>in</strong>crement <strong>in</strong> cGMP-IR<br />

when SNP stimulation was accompanied by the sGC<br />

<strong>in</strong>hibitor, ODQ (Fig. 4d). When ODQ was used alone,<br />

hardly any cGMP-IR became detectable (Fig. 4e). All these<br />

experiments showed the expression <strong>of</strong> a NO-sensitive sGC <strong>in</strong><br />

cells <strong>of</strong> the NT2 sphere culture.<br />

Fig. 5 The migration <strong>of</strong> cells out <strong>of</strong> NT2<br />

spheres was blocked <strong>in</strong> the presence <strong>of</strong><br />

NOS (7NI), sGC (ODQ), and PKG (RP-8-<br />

Br-cGMP) <strong>in</strong>hibitors. Migration was determ<strong>in</strong>ed<br />

after 24 h <strong>of</strong> exposure to chemical<br />

<strong>in</strong>hibitors. Representative photomicrographs<br />

<strong>of</strong> NT2 spheres <strong>in</strong>cubated as control<br />

(Ctrl) only with RA and 0.25% dimethylsulfoxide<br />

(DMSO) conta<strong>in</strong><strong>in</strong>g medium (a),<br />

500 lM 7NI (b), and 50 lM ODQ (c). Dosedependent<br />

<strong>in</strong>hibition <strong>of</strong> cell migration <strong>in</strong> the<br />

presence <strong>of</strong> 7NI (d), ODQ (e), and RP-8-BrcGMP<br />

(f). Scale bar: 100 lm. RP-8-BrcGMP,<br />

RP isomer <strong>of</strong> 8-Br-cGMP (**p <<br />

0.01, ***p < 0.001 compared to control.).<br />

NO is a positive regulator <strong>of</strong> cell migration via the cGMP<br />

and PKG pathway<br />

As NO/cGMP signal transduction could be a positive<br />

regulator <strong>of</strong> cell motility dur<strong>in</strong>g neural development, we<br />

exam<strong>in</strong>ed the effect <strong>of</strong> nNOS <strong>in</strong>hibition on migratory<br />

behavior <strong>of</strong> the differentiat<strong>in</strong>g NT2 cells. In the presence <strong>of</strong><br />

nNOS <strong>in</strong>hibitor, 7NI, the migration <strong>of</strong> cells out <strong>of</strong> the sphere<br />

was significantly reduced (Fig. 5a, b and d). Application <strong>of</strong><br />

the sGC <strong>in</strong>hibitor, ODQ, at a concentration <strong>of</strong> 200 lM<br />

decreased the migration <strong>of</strong> the cells by 60%. In a concentration<br />

range <strong>of</strong> 50–200 lM, ODQ <strong>in</strong>hibited cell migration<br />

<strong>in</strong> a dose-dependent manner (Fig. 5c and e). Potential<br />

(a) (b)<br />

(c) (d)<br />

(e) (f)<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841<br />

NO signal<strong>in</strong>g regulates neuronal precursor motility | 1835


1836 | M. A. Tegenge and G. Bicker<br />

downstream effector prote<strong>in</strong>s for the cGMP signal<strong>in</strong>g<br />

pathway are PKG that exist <strong>in</strong> two is<strong>of</strong>orms, PKG-I and<br />

PKG-II (Madhusoodanan and Murad 2007). The irreversibly<br />

PKG-I b<strong>in</strong>d<strong>in</strong>g <strong>in</strong>hibitor RP-8-Br-cGMP blocked the migration<br />

<strong>of</strong> cells <strong>in</strong> a dose-dependent manner (Fig. 5f).<br />

To test whether cell migration could be enhanced by an<br />

exogenous stimulation <strong>of</strong> NO/cGMP signal<strong>in</strong>g, we applied<br />

the NO donor NOC-18 which decays with long half-life<br />

(about 57 h at 22°C). After 24 h exposure to 50 lM <strong>of</strong> NOC-<br />

18, the migration <strong>of</strong> cells <strong>in</strong>creased by more than tw<strong>of</strong>old.<br />

Furthermore, NOC-18 significantly enhanced the migration<br />

<strong>in</strong> a dose-dependent manner <strong>in</strong> the concentration range <strong>of</strong> 1–<br />

100 lM (Fig. 6a–c). At relatively higher concentration<br />

(1 mM), NOC-18 appeared to significantly <strong>in</strong>hibit cell<br />

migration (Fig. 6c). A viability assay and morphological<br />

observation showed that 1 mM <strong>of</strong> NOC-18 significantly<br />

affected cell survival (Fig. 3e). The <strong>in</strong>hibitory effect on cell<br />

migration at this high concentration seemed to be because <strong>of</strong><br />

cytotoxicity (Figs. 3e and 6c). As we determ<strong>in</strong>ed cell<br />

migration based on the furthest migrated cells, an <strong>in</strong>creased<br />

rate <strong>of</strong> cell proliferation might contribute to the facilitation <strong>of</strong><br />

migration. Hence, we conducted a BrdU cell proliferation<br />

assay (Fig. 6d–h) <strong>in</strong> addition to viability test<strong>in</strong>g. At the<br />

concentration range where we found an <strong>in</strong>creas<strong>in</strong>g trend <strong>of</strong><br />

cell migration (1–50 lM), NOC-18 seemed to have a modest<br />

but not significant effect on cell proliferation (Fig. 6d–g).<br />

Intrigu<strong>in</strong>gly, 100 lM <strong>of</strong> NOC-18 appeared to facilitate cell<br />

migration while <strong>in</strong>hibit<strong>in</strong>g proliferation (Fig. 6c,f and g). A<br />

BrdU proliferation assay was also performed <strong>in</strong> the presence<br />

<strong>of</strong> nNOS/sGC/PKG <strong>in</strong>hibitors. We could not observe any<br />

effect on proliferation <strong>of</strong> NT2 cell spheres at the concentration<br />

<strong>of</strong> chemical <strong>in</strong>hibitors that blocked cell motility<br />

(Fig. 6h). Additionally, migration was determ<strong>in</strong>ed as the<br />

distance <strong>of</strong> the furthest migrated nest<strong>in</strong>-positive cells after<br />

NOC-18 treatment. We found out a significantly higher<br />

migration distance <strong>of</strong> the nest<strong>in</strong>-positive cells after 50 lM<br />

NOC-18 applications (Fig. 7a–c) suggest<strong>in</strong>g that NOC-18<br />

facilitated the migration <strong>of</strong> neuronal precursor cells. A BrdU<br />

proliferation assay was also performed for the nest<strong>in</strong>-positive<br />

cells which further confirmed that 50 lM NOC-18 did not<br />

alter the neuronal precursor cell proliferation (Fig. 7d).<br />

To demonstrate that the enhanced migration by exogenous<br />

NO application was mediated via the cGMP pathway<br />

(Fig. 3a), we exposed the NT2 cell spheres to NOC-18 while<br />

block<strong>in</strong>g the downstream target enzymes sGC and PKG. We<br />

observed a significant reduction <strong>of</strong> NO-<strong>in</strong>duced facilitation <strong>of</strong><br />

cell migration when NOC-18 was used <strong>in</strong> comb<strong>in</strong>ation with<br />

ODQ or RP-8-Br-cGMP (Fig. 8a). This provided direct<br />

evidence for the <strong>in</strong>volvement <strong>of</strong> cGMP and PKG <strong>in</strong> NOmediated<br />

neuronal precursor cell migration. Furthermore, a<br />

direct facilitation <strong>of</strong> cell migration was observed by a cell<br />

permeable analog <strong>of</strong> cGMP (8-Br-cGMP) (Fig. 8b–d). A<br />

rescue experiment showed that 8-Br-cGMP reversed the<br />

<strong>in</strong>hibitory effect <strong>of</strong> ODQ to the control level (Fig. 8e).<br />

Discussion<br />

NT2 cell spheres as a model for early develop<strong>in</strong>g human<br />

neuron<br />

The Ntera2 clone, D1 (NT2), is a well characterized cell l<strong>in</strong>e<br />

that has been derived from a human testicular cancer. NT2<br />

cells term<strong>in</strong>ally differentiate <strong>in</strong>to post-mitotic neurons by<br />

exposure to micromolar doses <strong>of</strong> RA and mitotic <strong>in</strong>hibitors<br />

(Andrews 1984; Pleasure et al. 1992; Paquet-Durand et al.<br />

2003). The differentiated neurons express a large number <strong>of</strong><br />

neuronal markers and form functional synapses [for review<br />

see Paquet-Durand and Bicker (2007)].<br />

Neurosphere-based cultures have been employed to study<br />

the <strong>in</strong>volvement <strong>of</strong> signal<strong>in</strong>g pathways that regulate neuronal<br />

precursor cell proliferation, differentiation, and migration<br />

(Leone et al. 2005; Mizuno et al. 2005; Moors et al. 2007).<br />

Here, we used 10 lM <strong>of</strong> RA to generate NT2 aggregate<br />

cultures that expresses both neuronal progenitor and early<br />

neuronal markers. When the cultures were seeded on PDL<br />

and Matrigel-coated substrate, cells could migrate out <strong>of</strong> the<br />

spheres. Most <strong>of</strong> the cells that migrated out after 1 week RA<br />

treatment were nest<strong>in</strong>-positive with only few cells display<strong>in</strong>g<br />

bIII-tubul<strong>in</strong> sta<strong>in</strong><strong>in</strong>g. This <strong>in</strong>dicates that the migrat<strong>in</strong>g cell<br />

population is ma<strong>in</strong>ly composed <strong>of</strong> neuronal precursor cells<br />

(Fig. 2a and c). When the time <strong>of</strong> RA application was<br />

prolonged to 2 week, we observed an <strong>in</strong>crease <strong>in</strong> the level <strong>of</strong><br />

bIII-tubul<strong>in</strong> <strong>in</strong> cell somata and the neuronal processes.<br />

Expression <strong>of</strong> neuronal markers such as glutamate receptors,<br />

MAP2, Tau, and NeuN just after 14-day exposure with RA<br />

has been reported previously for NT2 cells (Megiorni et al.<br />

2005). Furthermore, the migration assay showed that upon<br />

prolonged treatment with RA, the cells tended to rema<strong>in</strong><br />

<strong>in</strong>side the sphere and send out neuronal processes. This<br />

<strong>in</strong>dicates that RA treatment <strong>in</strong>duces neuronal differentiation<br />

while decreas<strong>in</strong>g cell migration (Figs. 1 and 2). RA has been<br />

shown to reduce the migration <strong>of</strong> airway smooth muscle cells<br />

(Day et al. 2006) and neuroblastoma cell l<strong>in</strong>es (Voigt and<br />

Z<strong>in</strong>tl 2003; Messi et al. 2008). The early expression <strong>of</strong><br />

cytoskeletal neuronal markers that resemble mammalian<br />

neurogenesis (Przyborski et al. 2000, 2003; Houldsworth<br />

et al. 2002; Paquet-Durand and Bicker 2007) suggests that<br />

the NT2 aggregate culture system is a valid model for<br />

develop<strong>in</strong>g human neuronal cells.<br />

NO/cGMP/PKG pathway mediates the migration <strong>of</strong> human<br />

neuronal precursor cells<br />

In this study we showed for the first time that NO signal<strong>in</strong>g<br />

positively regulated cell migration <strong>in</strong> a model <strong>of</strong> human<br />

neuronal precursor cells. Both undifferentiated NT2 precursors<br />

and differentiat<strong>in</strong>g NT2 cell spheres expressed nNOS<br />

(Fig. 4a) that could serve as endogenous source <strong>of</strong> NO. The<br />

presence <strong>of</strong> functional sGC was directly demonstrated by<br />

exogenous stimulation <strong>of</strong> cells with a NO donor and<br />

subsequent immunocytochemical detection <strong>of</strong> cGMP. The<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841


(a) (b) (c)<br />

(d) (e)<br />

(g) (h)<br />

Fig. 6 The NO donor (NOC-18) facilitates cell migration <strong>in</strong> a dosedependent<br />

manner. Migration distance was determ<strong>in</strong>ed after 24 h <strong>of</strong><br />

exposure to NOC-18. Representative photomicrographs <strong>of</strong> NT2 cell<br />

spheres under control condition (a) and exposed to 50 lM <strong>of</strong> NOC-18<br />

(b). Significant <strong>in</strong>crease <strong>in</strong> cell migration <strong>in</strong> the presence <strong>of</strong> NOC-18 at<br />

a concentration range <strong>of</strong> 1–100 lM (c). BrdU immun<strong>of</strong>luorescence<br />

(<strong>in</strong>tensity <strong>in</strong>verted) <strong>in</strong> a control (d), 50 lM (e), and 100 lM (f) <strong>of</strong> NOC-<br />

18. 4¢-6-diamid<strong>in</strong>o-2-phenyl<strong>in</strong>dole (DAPI) sta<strong>in</strong><strong>in</strong>g is shown at the<br />

NO-<strong>in</strong>duced cGMP-IR was reduced by the specific sGC<br />

<strong>in</strong>hibitor ODQ (Fig. 4b–e). After show<strong>in</strong>g the presence <strong>of</strong> the<br />

NO signal transduction pathway <strong>in</strong> the NT2 cultures we<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841<br />

NO signal<strong>in</strong>g regulates neuronal precursor motility | 1837<br />

(f)<br />

bottom <strong>of</strong> panels (d–f). NOC-18 did not affect cell proliferation <strong>in</strong> a<br />

concentration range <strong>of</strong> 1–50 lM; however, at 100 and 1000 lM, it<br />

significantly <strong>in</strong>hibits cell proliferation (g). Application <strong>of</strong> 7NI (500 lM),<br />

ODQ (50 lM), and RP-8-Br-cGMP (100 nM) to NT2 cell spheres did<br />

not alter cell proliferation (h). Scale bars: 100 lm. NOC-18, 2, 2-hydroxynitrosohydraz<strong>in</strong>o-bis-ethanam<strong>in</strong>e;<br />

RP-8-Br-cGMP, RP isomer <strong>of</strong><br />

8-Br-cGMP; Ctrl, control (*p < 0.05, **p < 0.01, ***p < 0.001 compared<br />

to control.).<br />

asked whether NO modulated the motility <strong>of</strong> cells out <strong>of</strong> the<br />

spheres. Chemical manipulation <strong>of</strong> the target enzymes<br />

<strong>in</strong>volved <strong>in</strong> NO transduction demonstrated that <strong>in</strong>hibition


1838 | M. A. Tegenge and G. Bicker<br />

(c)<br />

(d)<br />

(a) (b)<br />

Fig. 7 NOC-18 facilitates the migration <strong>of</strong> nest<strong>in</strong>-positive precursor<br />

cells. Migration was determ<strong>in</strong>ed after 48 h <strong>of</strong> exposure to NOC-18,<br />

fixed and sta<strong>in</strong>ed for nest<strong>in</strong>. Representative photomicrographs <strong>of</strong> NT2<br />

cell spheres under control condition (a) and exposed to 50 lM <strong>of</strong><br />

NOC-18 (b). The white l<strong>in</strong>es <strong>in</strong> a and b del<strong>in</strong>eate distance <strong>of</strong> migration<br />

from the rim <strong>of</strong> the sphere to the furthest migrated nest<strong>in</strong>-positive cells.<br />

Fifty lM <strong>of</strong> NOC-18 significantly <strong>in</strong>creased the migration <strong>of</strong> the nest<strong>in</strong>positive<br />

neuronal precursor cells (c) and the % BrdU sta<strong>in</strong><strong>in</strong>g <strong>of</strong> the<br />

nest<strong>in</strong>-positive cells was not altered by NOC-18 application (d). NOC-<br />

18, 2, 2-hydroxynitrosohydraz<strong>in</strong>o-bis-ethanam<strong>in</strong>e; Ctrl, control (***p <<br />

0.001).<br />

<strong>of</strong> nNOS, sGC, and PKG reduced the motility <strong>of</strong> cells<br />

(Fig. 5a–f). These loss <strong>of</strong> function effects suggested that NO<br />

signal transduction positively regulated cell motility. Possible<br />

cytotoxic effects <strong>of</strong> the <strong>in</strong>hibitors could be excluded as the<br />

concentrations we used were well below the dose that caused<br />

significant cell death (Fig. 3b–f).<br />

A second l<strong>in</strong>e <strong>of</strong> evidence derives from a ga<strong>in</strong> <strong>of</strong> function<br />

after the exogenous application <strong>of</strong> NO and cGMP. The NO<br />

donor and membrane permeable analog <strong>of</strong> cGMP enhanced<br />

cell migration (Figs. 6a–c,7a–c and 8b–d) which directly<br />

confirmed the positive regulatory role <strong>of</strong> NO and cGMP.<br />

Interest<strong>in</strong>gly, NO-<strong>in</strong>duced facilitation <strong>of</strong> cell migration was<br />

abolished when NOC-18 was used together with the sGC<br />

<strong>in</strong>hibitor ODQ or PKG <strong>in</strong>hibitor RP-8-Br-cGMP (Fig. 8a).<br />

This <strong>in</strong>hibition <strong>of</strong> migration by sGC and PKG <strong>in</strong>hibitors<br />

<strong>in</strong>fers that a certa<strong>in</strong> level <strong>of</strong> cGMP is required for cell<br />

motility. As pharmacological <strong>in</strong>hibition <strong>of</strong> the sGC enzyme<br />

can be completely rescued by application <strong>of</strong> the cGMP<br />

analog (Fig. 8e), it is unlikely that unspecific side effects <strong>of</strong><br />

the chemical blocker contribute to <strong>in</strong>hibition <strong>of</strong> cell migration.<br />

Our comb<strong>in</strong>ed result implicate that NO facilitates cell<br />

migration by <strong>in</strong>duc<strong>in</strong>g sGC to <strong>in</strong>crease cGMP levels which <strong>in</strong><br />

turn may activate PKG.<br />

NO signal transduction and neuronal motility<br />

<strong>Nitric</strong> oxide and cGMP have been shown to mediate the<br />

migration <strong>of</strong> various cell types <strong>in</strong>clud<strong>in</strong>g neutrophils,<br />

epithelial, and endothelial cells (Ziche et al. 1994; Elfer<strong>in</strong>k<br />

and VanUffelen 1996; Noiri et al. 1996, 1998). Chemical<br />

manipulations <strong>of</strong> cultured nervous systems implicate NO/<br />

cGMP as regulator also for neuronal cell motility [reviewed<br />

by Bicker (2005, 2007)]. Neuroanatomical studies us<strong>in</strong>g<br />

markers aga<strong>in</strong>st NOS and sGC suggest the migrat<strong>in</strong>g<br />

neuroblasts <strong>of</strong> the rostral migratory stream as potential<br />

targets for NO signal<strong>in</strong>g <strong>in</strong> the adult bra<strong>in</strong> (Moreno-López<br />

et al. 2000; Gutièrrez-Mec<strong>in</strong>as et al. 2007). NO signal<br />

transduction pathways regulate the migration <strong>of</strong> cerebellar<br />

neurons, <strong>in</strong>sect enteric neurons, and early develop<strong>in</strong>g<br />

Xenopus neuronal cells (Tanaka et al. 1994; Haase and<br />

Bicker 2003; Peunova et al. 2007; Knipp and Bicker 2009).<br />

The application <strong>of</strong> a NO donor to young adult rats before<br />

and after stroke <strong>in</strong>creased cell proliferation and migration <strong>in</strong><br />

the SVZ and dentate gyrus (Zhang et al. 2001). Some<br />

studies report that NO has no effect on post-mitotic<br />

neuronal cell migration (Moreno-López et al. 2004) or<br />

decreased ependymal/SVZ cell migration after focal cerebral<br />

ischemia <strong>in</strong> rats (Zhang et al. 2007). Such <strong>in</strong>consistency<br />

<strong>in</strong> the effects <strong>of</strong> NO on neuronal migration could be<br />

because <strong>of</strong> the use <strong>of</strong> different drug concentrations, species<br />

differences, pathological, and developmental stages <strong>of</strong> the<br />

experimental animals. Here, we provide further evidence for<br />

a significant role <strong>of</strong> NO <strong>in</strong> human neuronal precursor cell<br />

migration. The present study shows a concentration-dependent<br />

facilitation <strong>of</strong> human neuronal precursor cell migration<br />

by NO donor application. The <strong>in</strong>verted U-shaped dose-–<br />

response curve (Fig. 6c) might <strong>in</strong> part account for the<br />

<strong>in</strong>consistency <strong>in</strong> the role <strong>of</strong> NO. In our cell culture model,<br />

NOC-18 <strong>in</strong>creased the migration <strong>of</strong> cells out <strong>of</strong> spheres <strong>in</strong> a<br />

concentration range <strong>of</strong> 1–50 lM without affect<strong>in</strong>g cell<br />

viability and proliferation (Figs. 3e and 6a–g). Analysis <strong>of</strong><br />

migration distance and proliferation <strong>of</strong> the nest<strong>in</strong>-positive<br />

cells confirmed that NOC-18 facilitated the migration <strong>of</strong> the<br />

neuronal precursor cells (Fig. 7a–c) without affect<strong>in</strong>g<br />

proliferation (Fig. 7d). At 100 lM NOC-18 cell migration<br />

was enhanced while proliferation was reduced (Fig. 6c, f<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841


Fig. 8 NO facilitates migration <strong>of</strong> cells out<br />

<strong>of</strong> spheres via the cGMP and PKG pathway.<br />

NO-<strong>in</strong>duced facilitation <strong>of</strong> migration<br />

was blocked <strong>in</strong> the presence <strong>of</strong> sGC (50 lM<br />

ODQ) and PKG (100 nM RP-8-Br-cGMP)<br />

<strong>in</strong>hibitors (a). Direct facilitation <strong>of</strong> cell<br />

migration by cGMP analog (8-Br-cGMP) (b–<br />

d). Inhibitory effect <strong>of</strong> ODQ (50 lM) was<br />

rescued to the control level with 8-Br-cGMP<br />

(1 mM) (e). RP-8-Br-cGMP, RP isomer <strong>of</strong><br />

8-Br-cGMP; Ctrl, control (**p < 0.01, ***p <<br />

0.001 compared to control.)<br />

and g). This result implicates that NO differentially<br />

coord<strong>in</strong>ates cell migration and proliferation, as has also<br />

been reported by Peunova et al. (2007). At the relatively<br />

high concentration <strong>of</strong> 1 mM, the <strong>in</strong>crement <strong>in</strong> cell migration<br />

was not observed. This reduction was most probably<br />

because <strong>of</strong> cytotoxicity (Figs. 3e and 6g). Low concentration<br />

<strong>of</strong> exogenous NO has been shown to enhance random<br />

migration <strong>of</strong> neutrophils by stimulat<strong>in</strong>g sGC, while a higher<br />

concentration <strong>in</strong>hibited migration (Elfer<strong>in</strong>k and VanUffelen<br />

1996). Presumably, at low concentration NO activates only<br />

its ma<strong>in</strong> target enzyme sGC (Garthwaite 2008) whereas at<br />

sufficiently higher concentrations it rapidly reacts with<br />

prote<strong>in</strong>s and may even cause cell death [for review see<br />

(a)<br />

(b) (c)<br />

(d) (e)<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841<br />

NO signal<strong>in</strong>g regulates neuronal precursor motility | 1839<br />

Blaise et al. (2005) and Madhusoodanan and Murad<br />

(2007)].<br />

Presently, we do not know how the NO/cGMP/PKG<br />

pathway causes the cytoskeletal reorganization that leads to<br />

cell migration out <strong>of</strong> the NT2 spheres. In primary cultured<br />

smooth muscle cells, the migration stimulatory effect <strong>of</strong><br />

NO donors and cGMP is associated with altered cell<br />

morphology and dissociation <strong>of</strong> act<strong>in</strong> filaments (Brown<br />

et al. 1999). In <strong>in</strong>sect enteric neurons, where the NO/<br />

cGMP/PKG pathway is crucial for cell migration, a block<br />

<strong>of</strong> this pathway results <strong>in</strong> a relocalization <strong>of</strong> F-act<strong>in</strong><br />

bundles from the neurites to the cell body (Haase and<br />

Bicker 2003). The stimulation <strong>of</strong> the NO/cGMP/PKG


1840 | M. A. Tegenge and G. Bicker<br />

pathway <strong>in</strong> astrocytes caused a redistributation <strong>of</strong> glial<br />

fibrillary acidic prote<strong>in</strong> and depolymerization <strong>of</strong> act<strong>in</strong> with<br />

a loss <strong>of</strong> stress fibers. These cytoskeletal rearrangements<br />

resulted <strong>in</strong> a facilitation <strong>of</strong> astrocyte motility (Borán and<br />

García 2007). Such changes <strong>in</strong> neur<strong>of</strong>ilaments and act<strong>in</strong> are<br />

attributed to <strong>in</strong>hibition <strong>of</strong> RhoA GTPase by NO-stimulated<br />

PKG pathway (Sawada et al. 2001; Gudi et al. 2002; Borán<br />

and García 2007; Peunova et al. 2007). Thus, there is a<br />

l<strong>in</strong>k between the NO signal transduction pathway and act<strong>in</strong><br />

cytoskeleton, possibly via RhoA GTPase. Phosphorylation<br />

<strong>of</strong> Enabled/vasodilator-stimulated phosphoprote<strong>in</strong> family<br />

prote<strong>in</strong>s that regulate act<strong>in</strong> polymerization by PKG has<br />

also been reported to account for the action <strong>of</strong> cGMP <strong>in</strong><br />

cell motility (Sporbert et al. 1999; Chen et al. 2007;<br />

L<strong>in</strong>dsay et al. 2007). Most likely changes <strong>in</strong> the cytosolic<br />

calcium concentration will contribute to the morphology <strong>of</strong><br />

motile cell processes. Recent <strong>in</strong>vestigations have shown<br />

that NO regulates growth cone filopodial behavior <strong>in</strong>volv<strong>in</strong>g<br />

sGC, PKG, and cyclic adenos<strong>in</strong>e diphosphate ribose which<br />

causes the release <strong>of</strong> calcium from <strong>in</strong>tracellular stores via<br />

the ryanod<strong>in</strong>e receptor (Welshhans and Rehder 2005, 2007).<br />

This study demonstrated that NO facilitated the migration<br />

<strong>of</strong> differentiat<strong>in</strong>g model neuronal cells through the cGMP<br />

and PKG signal<strong>in</strong>g pathway. In our accessible NT2 sphere<br />

culture system we can now <strong>in</strong>vestigate the mechanisms by<br />

which NO regulates human neuronal precursor motility.<br />

Acknowledgements<br />

We thank Dr. J. de Vente for his k<strong>in</strong>d gift <strong>of</strong> the cGMP antiserum,<br />

Dr M. Stern for the discussion and help dur<strong>in</strong>g the preparation <strong>of</strong> the<br />

manuscript, and S. Tan for technical support. M.A. Tegenge was<br />

supported by a Georg–Christoph–Lichtenberg scholarship from the<br />

M<strong>in</strong>istry for Science and Culture <strong>of</strong> Lower Saxony and G. Bicker<br />

was supported by DFG grant BI 262/16-1.<br />

Support<strong>in</strong>g Information<br />

Additional Support<strong>in</strong>g Information may be found <strong>in</strong> the onl<strong>in</strong>e<br />

version <strong>of</strong> this article:<br />

Figure S1. BrdU is <strong>in</strong>corporated <strong>in</strong>to the nest<strong>in</strong>-positive neuronal<br />

precursor cells.<br />

As a service to our authors and readers, this journal provides<br />

support<strong>in</strong>g <strong>in</strong>formation supplied by the authors. Such materials are<br />

peer-reviewed and may be re-organized for onl<strong>in</strong>e delivery, but are<br />

not copy-edited or typeset. Technical support issues aris<strong>in</strong>g from<br />

support<strong>in</strong>g <strong>in</strong>formation (other than miss<strong>in</strong>g files) should be<br />

addressed to the authors.<br />

References<br />

Andrews P. W. (1984) Ret<strong>in</strong>oic acid <strong>in</strong>duces neuronal differentiation <strong>of</strong> a<br />

cloned human embryonal carc<strong>in</strong>oma cell l<strong>in</strong>e <strong>in</strong> vitro. Dev. Biol.<br />

103, 285–293.<br />

Ayala R., Shu T. and Tsai L. (2007) Trekk<strong>in</strong>g across the bra<strong>in</strong>: the<br />

journey <strong>of</strong> neuronal migration. Cell 128, 29–43.<br />

Bicker G. (2005) STOP and GO with NO: nitric oxide as a regulator <strong>of</strong><br />

cell motility <strong>in</strong> simple bra<strong>in</strong>s. BioEssays 27, 495–505.<br />

Bicker G. (2007) Pharmacological approaches to nitric oxide signall<strong>in</strong>g<br />

dur<strong>in</strong>g neural development <strong>of</strong> locusts and other model <strong>in</strong>sects.<br />

Arch. Insect Biochem. Physiol. 64, 43–58.<br />

Blaise G. A., Gauv<strong>in</strong> D., Gangal M. and Authier S. (2005) <strong>Nitric</strong> oxide,<br />

cell signal<strong>in</strong>g and cell death. Toxicology 208, 177–192.<br />

Borán M. S. and García A. (2007) The cyclic GMP-prote<strong>in</strong> k<strong>in</strong>ase G<br />

pathway regulates cytoskeleton dynamics and motility <strong>in</strong> astrocytes.<br />

J. Neurochem. 102, 216–230.<br />

Brown C., Pan X. and Hassid A. (1999) <strong>Nitric</strong> oxide and C-type atrial<br />

natriuretic peptide stimulate primary aortic smooth muscle cell<br />

migration via a cGMP-dependent mechanism: relationship to<br />

micr<strong>of</strong>ilament dissociation and altered cell morphology. Circ. Res.<br />

84, 655–667.<br />

Cárdenas A., Moro M. A., Hurtado O., Leza J. C. and Lizasoa<strong>in</strong> I. (2005)<br />

Dual role <strong>of</strong> nitric oxide <strong>in</strong> adult neurogenesis. Bra<strong>in</strong> Res. Bra<strong>in</strong><br />

Res. Rev. 50, 1–6.<br />

Chen H., Lev<strong>in</strong>e Y. C., Golan D. E., Michel T. and L<strong>in</strong> A. J. (2007)<br />

ANP-<strong>in</strong>itiated cGMP pathways regulate VASP phosphorylation<br />

and angiogenesis <strong>in</strong> vascular endothelium. J. Biol. Chem. 283,<br />

4439–4447.<br />

Day R. M., Lee Y. H., Park A. and Suzuki Y. J. (2006) Ret<strong>in</strong>oic acid<br />

<strong>in</strong>hibits airway smooth muscle cell migration. Am. J. Respir. Cell<br />

Mol. Biol. 34, 695–703.<br />

Elfer<strong>in</strong>k J. G. and VanUffelen B. E. (1996) The role <strong>of</strong> cyclic nucleotides<br />

<strong>in</strong> neutrophil migration. Gen. Pharmacol. 27, 387–393.<br />

Enikolopov G., Banerji J. and Kuz<strong>in</strong> B. (1999) <strong>Nitric</strong> oxide and Drosophila<br />

development. Cell Death Differ. 6, 956–963.<br />

Foster J. A. and Phelps P. E. (2000) Neurons express<strong>in</strong>g NADPHdiaphorase<br />

<strong>in</strong> the develop<strong>in</strong>g human sp<strong>in</strong>al cord. J. Comp. Neurol.<br />

427, 417–427.<br />

Gage F. H. (2000) Mammalian neural stem cells. Science 287, 1433–<br />

1438.<br />

Garthwaite J. (2008) Concepts <strong>of</strong> neural nitric oxide-mediated transmission.<br />

Eur. J. Neurosci. 27, 2783–2802.<br />

Gudi T., Chen J. C., Casteel D. E., Seasholtz T. M., Boss G. R. and Pilz<br />

R. B. (2002) cGMP-dependent prote<strong>in</strong> k<strong>in</strong>ase <strong>in</strong>hibits serumresponse<br />

element-dependent transcription by <strong>in</strong>hibit<strong>in</strong>g rho activation<br />

and functions. J. Biol. Chem. 277, 37382–37393.<br />

Gutièrrez-Mec<strong>in</strong>as M., Crespo C., Blasco-Ibáñez J. M., Nácher J., Varea<br />

E. and Martínez-Guijarro F. J. (2007) Migrat<strong>in</strong>g neuroblasts <strong>of</strong> the<br />

rostral migratory stream are putative targets for the action <strong>of</strong> nitric<br />

oxide. Eur. J. Neurosci. 26, 392–402.<br />

Haase A. and Bicker G. (2003) <strong>Nitric</strong> oxide and cyclic nucleotides are<br />

regulators <strong>of</strong> neuronal migration <strong>in</strong> an <strong>in</strong>sect embryo. Development<br />

130, 3977–3987.<br />

Hara K., Yasuhara T., Maki M. et al. (2008) Neural progenitor NT2N<br />

cell l<strong>in</strong>es from teratocarc<strong>in</strong>oma for transplantation therapy <strong>in</strong><br />

stroke. Prog. Neurobiol. 85, 318–334.<br />

Hatten M. E. (1999) Central nervous system neuronal migration. Annu.<br />

Rev. Neurosci. 22, 511–539.<br />

Houldsworth J., Heath S. C., Bosl G. J., Studer L. and Chaganti R. S.<br />

(2002) Expression pr<strong>of</strong>il<strong>in</strong>g <strong>of</strong> l<strong>in</strong>eage differentiation <strong>in</strong> pluripotential<br />

human embryonal carc<strong>in</strong>oma cells. Cell Growth Differ. 13,<br />

257–264.<br />

Hyun D. H., Lee M., Hattori N., Kubo S., Mizuno Y., Halliwell B. and<br />

Jenner P. (2002) Effect <strong>of</strong> wild-type or mutant Park<strong>in</strong> on oxidative<br />

damage, nitric oxide, antioxidant defenses, and the proteasome.<br />

J. Biol. Chem. 277, 28572–28577.<br />

Jurado S., Sánchez-Prieto J. and Torres M. (2004) Elements <strong>of</strong> the nitric<br />

oxide/cGMP pathway expressed <strong>in</strong> cerebellar granule cells: biochemical<br />

and functional characterisation. Neurochem. Int. 45, 833–<br />

843.<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841


Kawauchi T. and Hosh<strong>in</strong>o M. (2008) Molecular pathways regulat<strong>in</strong>g<br />

cytoskeletal organization and morphological changes <strong>in</strong> migrat<strong>in</strong>g<br />

neurons. Dev. Neurosci. 30, 30–46.<br />

Knipp S. and Bicker G. (2009) Regulation <strong>of</strong> enteric neuron migration<br />

by the gaseous messenger molecules CO and NO. Development<br />

136, 85–93.<br />

Lambert de Rouvroit C. L. and G<strong>of</strong>f<strong>in</strong>et A. M. (2001) Neuronal<br />

migration. Mech. Dev. 105, 47–56.<br />

Lee M., Hyun D. H., Jenner P. and Halliwell B. (2001) Effect <strong>of</strong> proteasome<br />

<strong>in</strong>hibition on cellular oxidative damage, antioxidant defences<br />

and nitric oxide production. J. Neurochem. 78, 32–41.<br />

Leone D. P., Relvas J. B., Campos L. S., Hemmi S., Brakebusch C.,<br />

Fässler R., Ffrench-Constant C. and Suter U. (2005) Regulation <strong>of</strong><br />

neural progenitor proliferation and survival by beta1 <strong>in</strong>tegr<strong>in</strong>s.<br />

J. Cell Sci. 118, 2589–2599.<br />

L<strong>in</strong>dsay S. L., Ramsey S., Aitchison M., Renné T. and Evans T. J. (2007)<br />

Modulation <strong>of</strong> lamellipodial structure and dynamics by NO-dependent<br />

phosphorylation <strong>of</strong> VASP Ser239. J. Cell Sci. 120, 3011–3021.<br />

Madhusoodanan K. S. and Murad F. (2007) NO-cGMP signall<strong>in</strong>g and<br />

regenerative medic<strong>in</strong>e <strong>in</strong>volv<strong>in</strong>g stem cells. Neurochem. Res. 32,<br />

681–694.<br />

Megiorni F., Mora B., Indov<strong>in</strong>a P. and Mazzilli M. C. (2005) Expression<br />

<strong>of</strong> neuronal markers dur<strong>in</strong>g NTera2/cloneD1 differentiation by cell<br />

aggregation method. Neurosci. Lett. 373, 105–109.<br />

Messi E., Florian M., Caccia C., Zanisi M. and Maggi R. (2008) Ret<strong>in</strong>oic<br />

acid reduces human neuroblastoma cell migration and <strong>in</strong>vasiveness:<br />

effects on DCX, LIS1, neur<strong>of</strong>ilaments-68 and viment<strong>in</strong><br />

expression. BMC Cancer 8, 30.<br />

Mizuno N., Kokubu H., Sato M., Nishimura A., Yamauchi J., Kurose H.<br />

and Itoh H. (2005) G prote<strong>in</strong>-coupled receptor signal<strong>in</strong>g through<br />

Gq and JNK negatively regulates neural progenitor cell migration.<br />

Proc. Natl Acad. Sci. USA 102, 12365–12370.<br />

Moors M., Cl<strong>in</strong>e J. E., Abel J. and Fritsche E. (2007) ERK-dependent<br />

and <strong>in</strong>dependent pathways trigger human neural progenitor cell<br />

migration. Toxicol. Appl. Pharmacol. 221, 57–67.<br />

Moreno-López B., Noval J. A., González-Bonet L. G. and Estrada C.<br />

(2000) Morphological bases for a role <strong>of</strong> nitric oxide <strong>in</strong> adult<br />

neurogenesis. Bra<strong>in</strong> Res. 869, 244–250.<br />

Moreno-López B., Romero-Grimaldi C., Noval J. A., Murillo-Carretero<br />

M., Matarredona E. R. and Estrada C. (2004) <strong>Nitric</strong> oxide is a<br />

physiological <strong>in</strong>hibitor <strong>of</strong> neurogenesis <strong>in</strong> the adult mouse subventricular<br />

zone and olfactory bulb. J. Neurosci. 24, 85–95.<br />

Nadarajah B., Alifragis P., Wong R. O. L. and Parnavelas J. G. (2003)<br />

Neuronal migration <strong>in</strong> the develop<strong>in</strong>g cerebral cortex: observations<br />

based on real-time imag<strong>in</strong>g. Cereb. Cortex 13, 607–611.<br />

Nelson P. T., Kondziolka D., Wechsler L. et al. (2002) Clonal human<br />

(hNT) neuron grafts for stroke therapy: neuropathology <strong>in</strong> a patient<br />

27 months after implantation. Am. J. Pathol. 160, 1201–1206.<br />

Noiri E., Peresleni T., Srivastava N., Weber P., Bahou W. F., Peunova N.<br />

and Goligorsky M. S. (1996) <strong>Nitric</strong> oxide is necessary for a switch<br />

from stationary to locomot<strong>in</strong>g phenotype <strong>in</strong> epithelial cells. Am. J.<br />

Physiol. 270, C794–C802.<br />

Noiri E., Lee E., Testa J., Quigley J., Colflesh D., Keese C. R., Giaever I.<br />

and Goligorsky M. S. (1998) Podok<strong>in</strong>esis <strong>in</strong> endothelial cell<br />

migration: role <strong>of</strong> nitric oxide. Am. J. Physiol. 274, C236–C244.<br />

Paquet-Durand F. and Bicker G. (2007) <strong>Human</strong> model neurons <strong>in</strong> studies<br />

<strong>of</strong> bra<strong>in</strong> cell damage and neural repair. Curr. Mol. Med. 7, 541–<br />

554.<br />

Paquet-Durand F., Tan S. and Bicker G. (2003) Turn<strong>in</strong>g teratocarc<strong>in</strong>oma<br />

cells <strong>in</strong>to neurons: rapid differentiation <strong>of</strong> NT-2 cells <strong>in</strong> float<strong>in</strong>g<br />

spheres. Bra<strong>in</strong> Res. Dev. Bra<strong>in</strong> Res. 142, 161–167.<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 110, 1828–1841<br />

NO signal<strong>in</strong>g regulates neuronal precursor motility | 1841<br />

Peunova N., Sche<strong>in</strong>ker V., Ravi K. and Enikolopov G. (2007) <strong>Nitric</strong><br />

oxide coord<strong>in</strong>ates cell proliferation and cell movements dur<strong>in</strong>g<br />

early development <strong>of</strong> Xenopus. Cell Cycle 6, 3132–3144.<br />

Pleasure S. J., Page C. and Lee V. M. (1992) Pure, postmitotic, polarized<br />

human neurons derived from NTera 2 cells provide a system for<br />

express<strong>in</strong>g exogenous prote<strong>in</strong>s <strong>in</strong> term<strong>in</strong>ally differentiated neurons.<br />

J. Neurosci. 12, 1802–1815.<br />

Podrygajlo G., Tegenge M. A., Gierse A., Paquet-Durand F., Tan S.,<br />

Bicker G. and Stern M. (2009) Cellular phenotypes <strong>of</strong> human<br />

model neurons (NT2) after differentiation <strong>in</strong> aggregate culture. Cell<br />

Tissue Res. 336, 439–452.<br />

Przyborski S. A., Morton I. E., Wood A. and Andrews P. W. (2000)<br />

Developmental regulation <strong>of</strong> neurogenesis <strong>in</strong> the pluripotent<br />

human embryonal carc<strong>in</strong>oma cell l<strong>in</strong>e NTERA-2. Eur. J. Neurosci.<br />

12, 3521–3528.<br />

Przyborski S. A., Smith S. and Wood A. (2003) Transcriptional pr<strong>of</strong>il<strong>in</strong>g<br />

<strong>of</strong> neuronal differentiation by human embryonal carc<strong>in</strong>oma stem<br />

cells <strong>in</strong> vitro. Stem Cells 21, 459–471.<br />

Sawada N., Itoh H., Yamashita J. et al. (2001) cGMP-dependent prote<strong>in</strong><br />

k<strong>in</strong>ase phosphorylates and <strong>in</strong>activates RhoA. Biochem. Biophys.<br />

Res. Commun. 280, 798–805.<br />

Sporbert A., Mertsch K., Smolenski A., Hasel<strong>of</strong>f R. F., Schönfelder G.,<br />

Paul M., Ruth P., Walter U. and Blasig I. E. (1999) Phosphorylation<br />

<strong>of</strong> vasodilator-stimulated phosphoprote<strong>in</strong>: a consequence <strong>of</strong><br />

nitric oxide- and cGMP-mediated signal transduction <strong>in</strong> bra<strong>in</strong><br />

capillary endothelial cells and astrocytes. Bra<strong>in</strong> Res. Mol. Bra<strong>in</strong><br />

Res. 67, 258–266.<br />

Tanaka M., Yoshida S., Yano M. and Hanaoka F. (1994) Roles <strong>of</strong><br />

endogenous nitric oxide <strong>in</strong> cerebellar cortical development <strong>in</strong> slice<br />

cultures. Neuroreport 5, 2049–2052.<br />

Tanaka J., Marker<strong>in</strong>k-van Ittersum M., Ste<strong>in</strong>busch H. W. and De Vente J.<br />

(1997) <strong>Nitric</strong> oxide-mediated cGMP synthesis <strong>in</strong> oligodendrocytes<br />

<strong>in</strong> the develop<strong>in</strong>g rat bra<strong>in</strong>. Glia 19, 286–297.<br />

de Vente J., Ste<strong>in</strong>busch H. W. and Schipper J. (1987) A new approach to<br />

immunocytochemistry <strong>of</strong> 3¢, 5¢-cyclic guanos<strong>in</strong>e monophosphate:<br />

preparation, specificity, and <strong>in</strong>itial application <strong>of</strong> a new antiserum<br />

aga<strong>in</strong>st formaldehyde-fixed 3¢, 5¢-cyclic guanos<strong>in</strong>e monophosphate.<br />

Neuroscience 22, 361–373.<br />

Voigt A. and Z<strong>in</strong>tl F. (2003) Effects <strong>of</strong> ret<strong>in</strong>oic acid on proliferation,<br />

apoptosis, cytotoxicity, migration, and <strong>in</strong>vasion <strong>of</strong> neuroblastoma<br />

cells. Med. Pediatr. Oncol. 40, 205–213.<br />

Welshhans K. and Rehder V. (2005) Local activation <strong>of</strong> the nitric oxide/<br />

cyclic guanos<strong>in</strong>e monophosphate pathway <strong>in</strong> growth cones regulates<br />

filopodial length via prote<strong>in</strong> k<strong>in</strong>ase G, cyclic ADP ribose and<br />

<strong>in</strong>tracellular Ca2+ release. Eur. J. Neurosci. 22, 3006–3016.<br />

Welshhans K. and Rehder V. (2007) <strong>Nitric</strong> oxide regulates growth cone<br />

filopodial dynamics via ryanod<strong>in</strong>e receptor-mediated calcium<br />

release. Eur. J. Neurosci. 26, 1537–1547.<br />

Zhang R., Zhang L., Zhang Z., Wang Y., Lu M., Lapo<strong>in</strong>te M. and Chopp<br />

M. (2001) A nitric oxide donor <strong>in</strong>duces neurogenesis and reduces<br />

functional deficits after stroke <strong>in</strong> rats. Ann. Neurol. 50, 602–611.<br />

Zhang P., Liu Y., Li J., Kang Q., Tian Y., Chen X., Zhao J., Shi Q. and<br />

Song T. (2007) Decreased neuronal nitric oxide synthase expression<br />

and cell migration <strong>in</strong> the peri-<strong>in</strong>farction after focal cerebral<br />

ischemia <strong>in</strong> rats. Neuropathology 27, 347–354.<br />

Zheng J. Q. and Poo M. (2007) Calcium signal<strong>in</strong>g <strong>in</strong> neuronal motility.<br />

Annu. Rev. Cell Dev. Biol. 23, 375–404.<br />

Ziche M., Morbidelli L., Mas<strong>in</strong>i E., Amer<strong>in</strong>i S., Granger H. J., Maggi<br />

C. A., Geppetti P. and Ledda F. (1994) <strong>Nitric</strong> oxide mediates<br />

angiogenesis <strong>in</strong> vivo and endothelial cell growth and migration <strong>in</strong><br />

vitro promoted by substance P. J. Cl<strong>in</strong>. Invest. 94, 2036–2044.


JOURNAL OF NEUROCHEMISTRY | 2009 | 111 | 1434–1446 doi: 10.1111/j.1471-4159.2009.06421.x<br />

, ,<br />

*Division <strong>of</strong> Cell Biology, Institute <strong>of</strong> Physiology, University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e Hannover, Hannover, Germany<br />

Center for Systems Neuroscience (ZSN), Hannover, Germany<br />

Abstract<br />

The human Ntera2 (NT2) teratocarc<strong>in</strong>oma cell l<strong>in</strong>e can be<br />

<strong>in</strong>duced to differentiate <strong>in</strong>to post-mitotic neurons. Here, we<br />

report that the human NT2 neurons generated by a spherical<br />

aggregate cell culture method express <strong>in</strong>creas<strong>in</strong>g levels <strong>of</strong><br />

typical pre-synaptic prote<strong>in</strong>s (synaps<strong>in</strong> and synaptotagm<strong>in</strong> I)<br />

along the neurite depend<strong>in</strong>g on the length <strong>of</strong> <strong>in</strong> vitro culture.<br />

By employ<strong>in</strong>g an antibody directed aga<strong>in</strong>st the lum<strong>in</strong>al doma<strong>in</strong><br />

<strong>of</strong> synaptotagm<strong>in</strong> I and the fluorescent dye N-(3-triethylammoniumpropyl)-4-(4-(dibutylam<strong>in</strong>o)styryl)pyrid<strong>in</strong>iumdibromide,<br />

we show that depolarized NT2 neurons display<br />

calcium-dependent exo-endocytotic synaptic vesicle recycl<strong>in</strong>g.<br />

NT2 neurons express the neuronal is<strong>of</strong>orm <strong>of</strong> neuronal<br />

nitric oxide synthase and soluble guanylyl cyclase (sGC), the<br />

major receptor for nitric oxide (NO). We tested whether NO<br />

The gaseous messenger nitric oxide (NO) which is produced<br />

by the enzyme nitric oxide synthase has been extensively<br />

studied as a retrograde messenger that modulates synaptic<br />

plasticity and memory formation (Boehn<strong>in</strong>g and Snyder<br />

2003; Garthwaite 2008; Taqatqeh et al. 2009). The major<br />

physiological target <strong>of</strong> NO is the cylic guanos<strong>in</strong>e-monophosphate<br />

(cGMP)-synthesiz<strong>in</strong>g enzyme soluble guanylyl<br />

cyclase (sGC). Numerous studies <strong>in</strong>dicate that NO can<br />

<strong>in</strong>duce transmitter release at pre-synaptic term<strong>in</strong>als (Hawk<strong>in</strong>s<br />

et al. 1994; Arancio et al. 1996; Meffert et al. 1996; Sporns<br />

and Jenk<strong>in</strong>son 1997; Wildemann and Bicker 1999; Li et al.<br />

2002; Nickels et al. 2007).<br />

Mechanisms by which NO <strong>in</strong>duces vesicle release are<br />

gradually be<strong>in</strong>g discovered. Enhancement <strong>of</strong> transmitter<br />

release via the cGMP pathway which certa<strong>in</strong>ly <strong>in</strong>volves<br />

phosphorylation via prote<strong>in</strong> k<strong>in</strong>ase G (PKG) has been<br />

suggested as a major mechanism (Arancio et al. 1995,<br />

2001; Lu et al. 1999; Wildemann and Bicker 1999; Li et al.<br />

2004; Garthwaite 2008). Moreover, post-translational modification<br />

<strong>of</strong> ion channels and prote<strong>in</strong>s <strong>in</strong>volved <strong>in</strong> vesicle<br />

dock<strong>in</strong>g/fusion processes by S-nitrosylation has also been<br />

signal transduction modulates synaptic vesicle turnover <strong>in</strong><br />

human NT2 neurons. NO donors and cylic guanos<strong>in</strong>e-monophosphate<br />

analogs enhanced synaptic vesicle recycl<strong>in</strong>g while<br />

a sGC <strong>in</strong>hibitor blocked the effect <strong>of</strong> NO donors. Two NO<br />

donors, sodium nitroprusside, and and N-Ethyl-2-(1-ethyl-2hydroxy-2-nitrosohydraz<strong>in</strong>o)<br />

ethanam<strong>in</strong>e evoked vesicle exocytosis<br />

which was partially blocked by the sGC <strong>in</strong>hibitor. The<br />

activator <strong>of</strong> adenylyl cyclase, forskol<strong>in</strong>, and a cAMP analog<br />

<strong>in</strong>duced synaptic vesicle recycl<strong>in</strong>g and exocytosis via a parallel<br />

act<strong>in</strong>g prote<strong>in</strong> k<strong>in</strong>ase A pathway. Our data from NT2<br />

neurons suggest that NO/cyclic nucleotide signal<strong>in</strong>g pathways<br />

may facilitate neurotransmitter release <strong>in</strong> human bra<strong>in</strong> cells.<br />

Keywords: cyclic AMP, cyclic GMP, exocytosis, FM1-43,<br />

nitric oxide, NTera2.<br />

J. Neurochem. (2009) 111, 1434–1446.<br />

reported (Meffert et al. 1996; Ahern et al. 2002). The NO/<br />

cGMP pathway has been shown to regulate synaptic vesicle<br />

endocytosis and recycl<strong>in</strong>g by <strong>in</strong>creas<strong>in</strong>g pre-synaptic<br />

Received August 7, 2009; revised manuscript received September 16,<br />

2009; accepted September 16, 2009.<br />

Address correspondence and repr<strong>in</strong>t requests to Gerd Bicker, Division<br />

<strong>of</strong> Cell Biology, Institute <strong>of</strong> Physiology, University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e<br />

Hannover, Bisch<strong>of</strong>sholer Damm 15, D-30173 Hannover, Germany.<br />

E-mail: gerd.bicker@tiho-hannover.de<br />

Abbrevations used: 8-Br-cAMP, 8-Bromoadenos<strong>in</strong>e 3¢, 5¢-cyclic<br />

adenos<strong>in</strong>e monophosphate, sodium salt; cGMP, cylic guanos<strong>in</strong>e-monophosphate;<br />

cGMP-IR, cGMP-immunoreactive; DIV, days <strong>in</strong> vitro; FM1-<br />

43, N-(3-triethylammoniumpropyl)-4-(4-(dibutylam<strong>in</strong>o)styryl)pyrid<strong>in</strong>ium<br />

dibromide; FM1-43Fx, fixable analog <strong>of</strong> FM1-43; H-89, N-[2-((p-<br />

Bromoc<strong>in</strong>namyl) am<strong>in</strong>o) ethyl]-5-isoqu<strong>in</strong>ol<strong>in</strong>esulfonamide, 2HCl; KRH,<br />

Krebs-R<strong>in</strong>ger’s – HEPES; nNOS, neuronal nitric oxide synthase; NO,<br />

nitric oxide; NOC-12, N-Ethyl-2-(1-ethyl-2-hydroxy-2-nitrosohydraz<strong>in</strong>o)<br />

ethanam<strong>in</strong>e; NT2, Ntera2; ODQ, 1H-[1,2,4]-oxadiazolo[4,3-a]qu<strong>in</strong>oxal<strong>in</strong>-1-one;<br />

PBS, phosphate-buffered sal<strong>in</strong>e; PFA, 4%<br />

paraformaldehyde <strong>in</strong> PBS; PKA, prote<strong>in</strong> k<strong>in</strong>ase A;PKG, prote<strong>in</strong> k<strong>in</strong>ase<br />

G;Rp-cAMP, adenos<strong>in</strong>e 3¢,5¢-cyclic monophosphorothioate, Rp-isomer;sGC,<br />

soluble guanylyl cyclase;SNP, sodium nitroprusside.<br />

Ó 2009 The Authors<br />

1434 Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446


phosphatidyl<strong>in</strong>ositol 4, 5-bisphosphate (Micheva et al.<br />

2003). In addition to its role as a retrograde messenger,<br />

NO has been implicated to play a critical role <strong>in</strong> early<br />

neuronal development <strong>in</strong>clud<strong>in</strong>g cell proliferation, migration,<br />

and synaptogenesis (Enikolopov et al. 1999; Bicker 2005,<br />

2007). However, the limited availability <strong>of</strong> <strong>in</strong> vitro models<br />

for the develop<strong>in</strong>g human bra<strong>in</strong> has restricted the analysis <strong>of</strong><br />

potential roles for NO signals <strong>in</strong> vesicle turnover, a pathway<br />

that may also contribute to activity-dependent nerve cell<br />

development.<br />

In this study, we used the well-characterized human<br />

teratocarc<strong>in</strong>oma cell l<strong>in</strong>e Ntera2 (NT2) that can be <strong>in</strong>duced to<br />

differentiate <strong>in</strong>to neurons upon treatment with ret<strong>in</strong>oic acid as<br />

surface-attached adherent monolayer culture (Andrews 1984;<br />

Pleasure et al. 1992). In recent years, this rather lengthy<br />

differentiation method was significantly reduced by employ<strong>in</strong>g<br />

a cell aggregate culture method (Paquet-Durand et al.<br />

2003; Podrygajlo et al. 2009). Despite their common clonal<br />

orig<strong>in</strong>, differentiated NT2 neurons comprise a heterogeneous<br />

population <strong>of</strong> cells express<strong>in</strong>g several neurotransmitters <strong>in</strong><br />

vitro (Guillema<strong>in</strong> et al. 2000; Podrygajlo et al. 2009).<br />

Electron microscopical and immunocytochemical studies<br />

revealed the presence <strong>of</strong> pre-synaptic vesicles and synaptic<br />

vesicle-associated prote<strong>in</strong>s such as synaptobrev<strong>in</strong>, synaptophys<strong>in</strong>,<br />

and synaps<strong>in</strong> (Sheridan and Maltese 1998; Hartley<br />

et al. 1999; Guillema<strong>in</strong> et al. 2000; Podrygajlo et al. 2009).<br />

NT2 neurons also express voltage-gated Na channels<br />

(Matsuoka et al. 1997) and L, N, P/Q and R-type Ca<br />

channels that are recruited for neurotransmission (Gao et al.<br />

1998; Neelands et al. 2000).<br />

Here, we characterized <strong>in</strong> aggregate culture generated<br />

human NT2 neurons by monitor<strong>in</strong>g lum<strong>in</strong>al synaptotagm<strong>in</strong> I<br />

immunoreactivity and imag<strong>in</strong>g <strong>of</strong> synaptic vesicle recycl<strong>in</strong>g<br />

with the fluorescent dye N-(3-triethylammoniumpropyl)-4-<br />

(4-(dibutylam<strong>in</strong>o)styryl)pyrid<strong>in</strong>ium dibromide (FM1-43).<br />

Both methods revealed that depolarized NT2 neurons display<br />

calcium-dependent synaptic vesicle recycl<strong>in</strong>g. By employ<strong>in</strong>g<br />

immunocytochemical methods, we showed the expression <strong>of</strong><br />

neuronal nitric oxide synthase (nNOS) and functional sGC <strong>in</strong><br />

subpopulations <strong>of</strong> NT2 neurons. F<strong>in</strong>ally, we used the human<br />

NT2 neurons as a model to demonstrate the <strong>in</strong>volvement <strong>of</strong><br />

NO/cGMP and cAMP/prote<strong>in</strong> k<strong>in</strong>ase A (PKA) signal<br />

transduction <strong>in</strong> vesicle recycl<strong>in</strong>g and pre-synaptic vesicle<br />

exocytosis.<br />

Materials and methods<br />

The NO donor N-Ethyl-2-(1-ethyl-2-hydroxy-2-nitrosohydraz<strong>in</strong>o)<br />

ethanam<strong>in</strong>e (NOC-12), the PKA antagonists N-[2-((p-Bromoc<strong>in</strong>namyl)<br />

am<strong>in</strong>o) ethyl]-5-isoqu<strong>in</strong>ol<strong>in</strong>esulfonamide, 2HCl (H-89) and<br />

adenos<strong>in</strong>e 3¢, 5¢-cyclic monophosphorothioate, Rp-isomer, triethylammonium<br />

salt were purchased from Calbiochem (Darmstadt,<br />

Germany). 8-bromoguanos<strong>in</strong>e-3¢, 5¢-cyclic monophosphate, sodium<br />

salt and 8-bromoadenos<strong>in</strong>e 3¢, 5¢-cyclic monophosphate, sodium<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446<br />

NO and cyclic nucleotide modulates vesicle release | 1435<br />

salt (8-Br-cAMP) were purchased from Alexis Biochemicals<br />

(Lörrach, Germany). All other materials were obta<strong>in</strong>ed from<br />

Sigma (Taufkirchen, Germany) unless otherwise noted. Buffers<br />

with the follow<strong>in</strong>g compositions were prepared: phosphate-buffered<br />

sal<strong>in</strong>e (PBS; 10 mM sodium phosphate, 150 mM NaCl, pH<br />

7.4), Krebs-R<strong>in</strong>ger’s – HEPES (KRH) (<strong>in</strong> mM; 115 NaCl, 5 KCl,<br />

1 MgCl 2 · 6H 2O, 24 NaHC0 3, 2.5 CaCl 2.2H 2O, 25 Glucose, 25<br />

HEPES, pH 7.4), calcium free KRH that conta<strong>in</strong>s 2.5 mM EGTA<br />

<strong>in</strong>stead <strong>of</strong> CaCl 2 · 2H 2O (Ca 2+ free), KRH buffer that conta<strong>in</strong>s<br />

60 mM <strong>of</strong> KCl, corrected for osmolarity by reduction <strong>of</strong> NaCl<br />

(high K + ) and high K + without CaCl 2 · 2H 2O (Ca 2+ free high K + ).<br />

All chemicals were diluted <strong>in</strong> KRH at the follow<strong>in</strong>g f<strong>in</strong>al<br />

concentrations: sodium nitroprusside (SNP, 1 mM), NOC-12<br />

(100 lM), 1H-[1, 2, 4]-oxadiazolo [4, 3-a] qu<strong>in</strong>oxal<strong>in</strong>-1-one<br />

(ODQ, 50 lM), forskol<strong>in</strong> (50–100 lM), 8-bromoguanos<strong>in</strong>e-3¢, 5¢cyclic<br />

monophosphate, sodium salt (100–1000 lM), 8-Br-cAMP<br />

(100–1000 lM), adenos<strong>in</strong>e 3¢, 5¢-cyclic monophosphorothioate,<br />

Rp-isomer (Rp-cAMP) (10 lM) and H-89 (10 lM). 3-Isobutyl-1methylxanth<strong>in</strong>e,<br />

ODQ, and forskol<strong>in</strong> were prepared from stocks <strong>in</strong><br />

dimethylsulfoxide, which were further diluted <strong>in</strong> KRH to result <strong>in</strong><br />

a maximum concentration <strong>of</strong> 0.25% dimethylsulfoxide. Pathway<br />

<strong>in</strong>hibitors were pre-<strong>in</strong>cubated for 30 m<strong>in</strong> with the NT2 neurons<br />

prior to the experiments while activators were applied for 5 m<strong>in</strong><br />

dur<strong>in</strong>g load<strong>in</strong>g <strong>of</strong> FM1-43.<br />

Cell culture<br />

The human NT2/D1 cell l<strong>in</strong>e was obta<strong>in</strong>ed from American Type<br />

Culture Collection, Manassas, VA, USA. NT2 precursor cells were<br />

ma<strong>in</strong>ta<strong>in</strong>ed and cultivated <strong>in</strong> Dulbecco’s modified Eagle medium/<br />

F12 (Gibco-Invitrogen, Karlsruhe, Germany) supplemented with<br />

10% fetal bov<strong>in</strong>e serum (Gibco-Invitrogen) and 1% penicill<strong>in</strong>/<br />

streptomyc<strong>in</strong> (Gibco-Invitrogen) <strong>in</strong> an atmosphere <strong>of</strong> 5% CO 2 at<br />

37°C (Andrews 1984). Post-mitotic NT2 neurons were obta<strong>in</strong>ed by<br />

free-float<strong>in</strong>g spherical aggregate methods (Paquet-Durand et al.<br />

2003; Podrygajlo et al. 2009). Briefly, NT2 precursor cells were<br />

seeded <strong>in</strong> 95 mm bacteriological grade Petri dishes (Gre<strong>in</strong>er,<br />

Hamburg, Germany) for 24 h. Dulbecco’s modified Eagle medium/<br />

F12 medium conta<strong>in</strong><strong>in</strong>g 10% fetal bov<strong>in</strong>e serum and 10 lM ret<strong>in</strong>oic<br />

acid was used up to 1 week. The aggregates were mechanically<br />

dispersed and treated <strong>in</strong> RA (ret<strong>in</strong>oic acid) for one additional week <strong>in</strong><br />

T75 flasks as adherent culture. F<strong>in</strong>ally, cells were tryps<strong>in</strong>ized and<br />

seeded <strong>in</strong> T75 flasks and supplied with culture medium conta<strong>in</strong><strong>in</strong>g<br />

mitotic <strong>in</strong>hibitors (1 lM 1-6-D-arab<strong>in</strong><strong>of</strong>uransylcytos<strong>in</strong>e, 10 lM 2¢deoxy-5-fluorourid<strong>in</strong>e,<br />

10 lM 1-b-D-rib<strong>of</strong>uranosyluracil). After 7–<br />

10 days, fully differentiated neurons were selectively tryps<strong>in</strong>ized and<br />

were plated on poly-D-lys<strong>in</strong>e and Matrigel (Becton-Dick<strong>in</strong>son,<br />

Bedford, MA, USA) coated 12 or 25 mm cover glasses at a density<br />

<strong>of</strong> 20 000 or 100 000 cells/cover glass for further experimentation.<br />

Immun<strong>of</strong>luorescence<br />

Ntera2 neurons were washed with PBS and fixed for 30 m<strong>in</strong> at<br />

20°C with 4% paraformaldehyde <strong>in</strong> PBS (PFA). Immun<strong>of</strong>luorescence<br />

detection <strong>of</strong> microtubule associated prote<strong>in</strong> 2 (MAP2)<br />

(1 : 1000; Sigma), Tau1 (1 : 500; Millipore International),<br />

synaps<strong>in</strong> (1 : 500; Synaptic Systems, Gött<strong>in</strong>gen, Germany), synaptotagm<strong>in</strong><br />

I directed to cytoplasmic doma<strong>in</strong> (1 : 10, mAb30;<br />

Developmental Studies Hybridoma Bank, Iowa, USA), and nNOS<br />

(1 : 200; Sigma) was performed as previously described (Tegenge


1436 | M. A. Tegenge et al.<br />

and Bicker 2009). For the detection <strong>of</strong> cGMP-immunoreactive<br />

(cGMP-IR) cells, NT2 neurons were pre-<strong>in</strong>cubated for 20 m<strong>in</strong> at<br />

20°C with 1 mM SNP as an NO donor, 20 lM 3-(50-Hydroxymethyl-20-furyl)-1-benzyl<br />

<strong>in</strong>dazole as an enhancer <strong>of</strong> NO-<strong>in</strong>duced<br />

activity <strong>of</strong> sGC, and 1 mM 3-isobutyl-1-methylxanth<strong>in</strong>e as a<br />

phosphodiesterase <strong>in</strong>hibitor <strong>in</strong> PBS. Cultures were washed once<br />

with PBS and fixed with 4% PFA for 30 m<strong>in</strong>. The polyclonal<br />

sheep cGMP antiserum (1 : 10 000; a k<strong>in</strong>d gift from Dr. J. de<br />

Vente, Maastricht University, the Netherlands) was used as primary<br />

antibody to detect the level <strong>of</strong> cGMP (Tegenge and Bicker 2009).<br />

Lum<strong>in</strong>al synaptotagm<strong>in</strong> I immunosta<strong>in</strong><strong>in</strong>g<br />

Ntera2 neurons were assayed for synaptic vesicle recycl<strong>in</strong>g us<strong>in</strong>g<br />

polyclonal antibody directed aga<strong>in</strong>st the lum<strong>in</strong>al doma<strong>in</strong> <strong>of</strong><br />

synaptotagm<strong>in</strong> I (Synaptic Systems). Cultures were <strong>in</strong>cubated for<br />

20 m<strong>in</strong> with the antibody diluted 1 : 100 <strong>in</strong> high K + . For<br />

comparison, <strong>in</strong>corporation <strong>of</strong> anti-lum<strong>in</strong>al synaptotagm<strong>in</strong> I was<br />

performed <strong>in</strong> normal KRH (basal), Ca 2+ free high K + , and SNP or<br />

SNP + ODQ diluted <strong>in</strong> KRH. Cells were then fixed with 4% PFA<br />

and the <strong>in</strong>corporated antibody was detected by immun<strong>of</strong>luorescence.<br />

Western blott<strong>in</strong>g<br />

The prote<strong>in</strong> content <strong>of</strong> cell lysate was estimated by the BCA TM<br />

prote<strong>in</strong> assay kit (Pierce, Rockford, IL, USA). Equal amount <strong>of</strong><br />

prote<strong>in</strong> (50 lg) from NT2 cells and NT2 neurons was used. The<br />

monoclonal anti-nNOS (Sigma) at 1 : 1000 dilution and antiacetylated-a-tubul<strong>in</strong><br />

(Sigma) diluted 1 : 10 000 were used for<br />

western blott<strong>in</strong>g as previously described (Tegenge and Bicker 2009).<br />

FM1-43 imag<strong>in</strong>g<br />

The uptake and release <strong>of</strong> FM1-43 dye was performed as described<br />

by Gaffield and Betz (2006). Briefly, for uptake experiments, NT2<br />

neurons cultured for 3–5 weeks were washed with PBS and<br />

stimulated for 5 m<strong>in</strong> with high K + buffer <strong>in</strong> the presence <strong>of</strong><br />

10 lM fixable analog <strong>of</strong> FM1-43 (FM1-43Fx; Molecular Probes,<br />

Eugene, OR, USA) and compared with the uptake <strong>of</strong> FM1-43 <strong>in</strong><br />

normal KRH (basal), Ca 2+ free high K + and chemicals. After FM1-<br />

43Fx was loaded <strong>in</strong>to pre-synaptic term<strong>in</strong>als, NT2 neurons were<br />

washed twice <strong>in</strong> KRH followed by additional twice wash<strong>in</strong>g <strong>in</strong> Ca 2+<br />

free buffer. Neurons were then fixed with 4% PFA and images were<br />

taken immediately. The release <strong>of</strong> normal FM1-43 was followed <strong>in</strong><br />

real time after load<strong>in</strong>g the NT2 neurons with 10 lM FM1-43<br />

(Molecular Probes) <strong>in</strong> high K + buffer for 5 m<strong>in</strong>. The culture was<br />

transferred <strong>in</strong>to a perfusion chamber, mounted on a Zeiss Axiovert<br />

200 microscope and ma<strong>in</strong>ta<strong>in</strong>ed at 37°C. Cultures were cont<strong>in</strong>uously<br />

washed for 10 m<strong>in</strong> <strong>in</strong> normal KRH <strong>in</strong> a perfusion chamber at<br />

flow rate <strong>of</strong> 2–2.5 mL/m<strong>in</strong>. Cultures were washed for additional<br />

5 m<strong>in</strong> <strong>in</strong> Ca 2+ free buffer. The unload<strong>in</strong>g <strong>of</strong> FM1-43 was followed<br />

upon stimulation <strong>of</strong> the culture with high K + ,Ca 2+ free high K + ,<br />

normal KRH, or chemicals diluted <strong>in</strong> normal KRH.<br />

Microscopy and data analysis<br />

Preparations were viewed with a Zeiss Axiovert 200 (Gött<strong>in</strong>gen,<br />

Germany), equipped with a CoolSnap camera (Photometrics,<br />

Tucson, AZ, USA) and MetaMorph s<strong>of</strong>tware (Molecular Devices,<br />

Sunnyvale, CA, USA). Confocal images <strong>of</strong> NT2 neurons loaded<br />

with FM1-43Fx and co-sta<strong>in</strong>ed aga<strong>in</strong>st cytoplasmic doma<strong>in</strong> <strong>of</strong><br />

synaptotagm<strong>in</strong> I were prepared us<strong>in</strong>g a Leica TCS-SP5 spectral laser<br />

scann<strong>in</strong>g confocal microscope (Leica Mikrosysteme Vertrieb<br />

GmbH, Wetzlar, Germany) with LAS AF s<strong>of</strong>tware and under 63·<br />

oil immersion objective. For quantification <strong>of</strong> synaps<strong>in</strong> and<br />

synaptotagm<strong>in</strong> levels, sta<strong>in</strong>ed synaptic puncta were counted along<br />

the length <strong>of</strong> a neurite and expressed as number <strong>of</strong> synaptic puncta<br />

per number <strong>of</strong> neurons obta<strong>in</strong>ed from 4¢,6-diamid<strong>in</strong>o-2¢-phenyl<strong>in</strong>dol-dihydrochloride<br />

sta<strong>in</strong><strong>in</strong>g. To obta<strong>in</strong> an estimate <strong>of</strong> cGMP levels,<br />

we counted cGMP-IR cell bodies. Data were presented as<br />

mean ± SEM <strong>of</strong> at least eight fields <strong>of</strong> view (220 · 165 lm) from<br />

at least three <strong>in</strong>dependent experiments. For FM1-43Fx uptake<br />

experiments, fluorescence <strong>in</strong>tensity <strong>of</strong> <strong>in</strong>dividual puncta (numbers<br />

are given <strong>in</strong> the figure legends) was measured us<strong>in</strong>g Simple PCI<br />

s<strong>of</strong>tware (C-Imag<strong>in</strong>g Systems, Sewickley, PA, USA). Background<br />

fluorescence from a selected blank area was subtracted from<br />

<strong>in</strong>dividual puncta. For FM1-43 unload<strong>in</strong>g experiments, fluorescence<br />

<strong>in</strong>tensity was measured as above every 15 s for about 5 m<strong>in</strong>. The<br />

percent fluorescence <strong>in</strong>tensity rema<strong>in</strong><strong>in</strong>g at the end <strong>of</strong> stimulation<br />

was calculated. Data were presented as mean ± SEM. Statistical<br />

analysis was performed us<strong>in</strong>g unpaired Student’s t-test. Levels <strong>of</strong><br />

significance were: *p < 0.05, **p < 0.01, ***p < 0.001.<br />

Results<br />

NT2 neurons undergo pre-synaptic maturation<br />

We used the aggregate culture method (Paquet-Durand<br />

et al. 2003; Podrygajlo et al. 2009) to generate NT2<br />

neurons express<strong>in</strong>g MAP2 and Tau, typical neuronal<br />

cytoskeletal markers <strong>of</strong> dendrites and axons, respectively.<br />

NT2 neurons cultured for 7 days <strong>in</strong> vitro (DIV) after<br />

differentiation displayed MAP2 sta<strong>in</strong><strong>in</strong>g on short processes<br />

orig<strong>in</strong>at<strong>in</strong>g from the cell somata (Fig. 1a). Only weak Tausta<strong>in</strong><strong>in</strong>g<br />

appeared on long-neuronal processes after 7 DIV<br />

(Fig. 1b). NT2 neurons cultured for 28 DIV were sta<strong>in</strong>ed<br />

strongly for both MAP2 and Tau (Fig. 1e and f). To<br />

monitor the progress <strong>of</strong> pre-synaptic maturation, fully<br />

differentiated NT2 neurons were cultured <strong>in</strong> vitro for about<br />

6 weeks and sta<strong>in</strong>ed over time for the pre-synaptic prote<strong>in</strong>s<br />

synaps<strong>in</strong> and synaptotagm<strong>in</strong> I. After 7 DIV culture, NT2<br />

neurons expressed very little punctate synaps<strong>in</strong> sta<strong>in</strong><strong>in</strong>g<br />

that appeared ma<strong>in</strong>ly around cell somata (Fig. 1c). However,<br />

with<strong>in</strong> 14–28 DIV <strong>in</strong>tense punctate synaps<strong>in</strong> sta<strong>in</strong><strong>in</strong>g<br />

appeared on long neurites and around cell somata<br />

(Fig. 1g). Similarly, only little punctate synaptotagm<strong>in</strong> I<br />

sta<strong>in</strong><strong>in</strong>g was conf<strong>in</strong>ed to the cell bodies <strong>of</strong> NT2 neurons<br />

cultured for 7 DIV (Fig. 1d). Intense synaptotagm<strong>in</strong> I<br />

sta<strong>in</strong><strong>in</strong>g appeared along the neurites with<strong>in</strong> 14–28 DIV<br />

cultures (Fig. 1h). We quantified the levels <strong>of</strong> synaps<strong>in</strong> and<br />

synaptotagm<strong>in</strong> I as number <strong>of</strong> synaptic puncta per neuron.<br />

The level <strong>of</strong> both prote<strong>in</strong>s <strong>in</strong>creased significantly with<strong>in</strong><br />

14–28 DIV culture as compared with 7 DIV (Fig. 2a)<br />

<strong>in</strong>dicat<strong>in</strong>g that human NT2 neurons undergo pre-synaptic<br />

maturation. In NT2 neurons cultured for about 6 weeks,<br />

the level <strong>of</strong> synaps<strong>in</strong> and synaptotagm<strong>in</strong> did not <strong>in</strong>crease<br />

compared with 28 DIV old cultures (data not shown).<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446


Fig. 1 <strong>Human</strong> NT2 neurons express typical<br />

cytoskeletal neuronal marker and presynaptic<br />

prote<strong>in</strong>s. The maturation <strong>of</strong> NT2<br />

neurons generated by aggregate culture<br />

was followed by immunocytochemical<br />

sta<strong>in</strong><strong>in</strong>g for (a, f) MAP2, (b, g) Tau, (c, h)<br />

synaps<strong>in</strong>, (d, i) cytoplasmic, and (e, j)<br />

lum<strong>in</strong>al synaptotagm<strong>in</strong> I. After 7 days<br />

<strong>in</strong> vitro (DIV), NT2 neurons were (a) <strong>in</strong>tensely<br />

sta<strong>in</strong>ed for the neuronal marker<br />

MAP2, (b) weakly sta<strong>in</strong>ed for the axonal<br />

marker, Tau. (c, d) Only little synaps<strong>in</strong> and<br />

synaptotagm<strong>in</strong> puncta sta<strong>in</strong><strong>in</strong>g appeared on<br />

NT2 neurons cultured for 7 DIV. After 28<br />

DIV culture, NT2 neurons were <strong>in</strong>tensely<br />

sta<strong>in</strong>ed for (f) MAP2 and (g) Tau. (h, i)<br />

Intense synaps<strong>in</strong> and synaptotagm<strong>in</strong> sta<strong>in</strong><strong>in</strong>g<br />

appeared along the neurites <strong>of</strong> NT2<br />

neurons cultured for 28 DIV. (e, j) Presynaptic<br />

term<strong>in</strong>als <strong>of</strong> NT2 neurons cultured<br />

for 28 DIV were labeled with lum<strong>in</strong>al<br />

synaptotagm<strong>in</strong> I compared with 7 DIV culture.<br />

Arrow heads (h, i and j) <strong>in</strong>dicate<br />

representative punctate sta<strong>in</strong><strong>in</strong>g along<br />

the neurites. Blue (4¢,6-diamid<strong>in</strong>o-2¢-phenyl<strong>in</strong>dol-dihydrochloride)<br />

<strong>in</strong>dicates nuclear<br />

counter-sta<strong>in</strong><strong>in</strong>g. Scale bar: 200 lm (a, b,<br />

f and g), 100 lm (c, d, h and i), 50 lm<br />

(e and j). [Correction added after onl<strong>in</strong>e<br />

publication 9 November 2009; panel labell<strong>in</strong>g<br />

changed <strong>in</strong> figure legend].<br />

(a) (f)<br />

(b) (g)<br />

(c) (h)<br />

(d) (i)<br />

(e) (j)<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446<br />

NO and cyclic nucleotide modulates vesicle release | 1437


1438 | M. A. Tegenge et al.<br />

(a)<br />

(b)<br />

(c)<br />

Fig. 2 <strong>Human</strong> NT2 neurons undergo pre-synaptic maturation. The<br />

levels <strong>of</strong> synaps<strong>in</strong> and synaptotagm<strong>in</strong> I were quantified by count<strong>in</strong>g the<br />

number <strong>of</strong> synaptic puncta along the neurites <strong>in</strong> at least eight fields <strong>of</strong><br />

view taken from at least three <strong>in</strong>dependent experiments. (a) The<br />

number <strong>of</strong> puncta sta<strong>in</strong><strong>in</strong>g for synaps<strong>in</strong> and synaptotagm<strong>in</strong> were significantly<br />

<strong>in</strong>creased with<strong>in</strong> 14–28 DIV culture as compared with 7 DIV.<br />

(b) Synaptic vesicle recycl<strong>in</strong>g was determ<strong>in</strong>ed by antibody directed to<br />

the lum<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> synaptotagm<strong>in</strong> I dur<strong>in</strong>g depolarization <strong>of</strong> human<br />

NT2 neurons by high K + . The extent <strong>of</strong> synaptic vesicle recycl<strong>in</strong>g <strong>in</strong>creased<br />

significantly dur<strong>in</strong>g the maturation <strong>of</strong> NT2 neurons with<strong>in</strong> 14–<br />

28 DIV. (c) The <strong>in</strong>corporation <strong>of</strong> lum<strong>in</strong>al synaptotagm<strong>in</strong> I depends on<br />

both depolarization and presence <strong>of</strong> calcium.<br />

Depolarized human NT2 neurons display synaptic vesicle<br />

recycl<strong>in</strong>g <strong>in</strong> a calcium-dependent manner<br />

Antibodies directed to the lum<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> synaptotagm<strong>in</strong><br />

I have been used to label pre-synaptic vesicles that undergo<br />

exocytosis (Matteoli et al. 1992; Kraszewski et al. 1995;<br />

Verderio et al. 2007). We applied the lum<strong>in</strong>al synaptotagm<strong>in</strong><br />

I antibody <strong>in</strong> the presence <strong>of</strong> a depolariz<strong>in</strong>g high K +<br />

(60 mM) which causes multiple round <strong>of</strong> synaptic vesicle<br />

recycl<strong>in</strong>g and hence <strong>in</strong>corporation <strong>of</strong> the antibody <strong>in</strong>to the<br />

nerve term<strong>in</strong>als. Upon stimulation with high K + , NT2<br />

neurons (28 DIV) were <strong>in</strong>tensely labeled with synaptotagm<strong>in</strong><br />

(Fig. 1j) <strong>in</strong>dicat<strong>in</strong>g that the neurons perform synaptic vesicle<br />

recycl<strong>in</strong>g. The immunoreactivity appeared as punctate sta<strong>in</strong><strong>in</strong>g<br />

along the neurites and cell somata (Fig. 1j). We found on<br />

average higher values <strong>of</strong> immunoreactive synaptic puncta <strong>in</strong><br />

NT2 neurons cultured for about 28 DIV than for 7 DIV<br />

(Figs 1i, j and 2b). Thus, similar to the expression <strong>of</strong> the presynaptic<br />

prote<strong>in</strong>s, the activity-dependent immunosta<strong>in</strong><strong>in</strong>g <strong>of</strong><br />

the neurons critically depended on the length <strong>of</strong> <strong>in</strong> vitro<br />

culture. Quantification <strong>of</strong> the number <strong>of</strong> synaptic puncta<br />

revealed that depolarization by high K + resulted <strong>in</strong> significantly<br />

higher values <strong>of</strong> lum<strong>in</strong>al synaptotagm<strong>in</strong> label<strong>in</strong>g<br />

compared with the basal level (Fig. 2c). Application <strong>of</strong> antisynaptotagm<strong>in</strong><br />

antibody under high K + stimulation <strong>in</strong> the<br />

absence <strong>of</strong> Ca 2+ resulted <strong>in</strong> significantly lowered numbers <strong>of</strong><br />

synaptic puncta (Fig. 2c). These results clearly demonstrate<br />

that depolarized NT2 neurons show Ca 2+ -dependent synaptic<br />

vesicle recycl<strong>in</strong>g.<br />

To directly visualize synaptic exo- and compensatory<br />

endocytosis, we used fluorescent FM-dyes that are trapped<br />

<strong>in</strong> retrieved vesicle membranes <strong>in</strong> an activity-dependent<br />

manner. Firstly, we performed vesicle uptake experiments<br />

with the fixable dye FM1-43Fx as a monitor <strong>of</strong> synaptic vesicle<br />

turnover (Fig. 3a–c). Upon stimulation with high K + , NT2<br />

neurons were successfully loaded with FM1-43Fx as compared<br />

with the basal level (Fig. 3a–c). Furthermore, the<br />

load<strong>in</strong>g <strong>of</strong> FM1-43Fx was significantly lowered when it was<br />

applied with high K + <strong>in</strong> the absence <strong>of</strong> Ca 2+ (Fig. 3c). Then,<br />

we imaged exocytosis by monitor<strong>in</strong>g the dimm<strong>in</strong>g <strong>of</strong> fluorescence<br />

when dye trapped <strong>in</strong> the synaptic vesicles is be<strong>in</strong>g<br />

released. The unload<strong>in</strong>g <strong>of</strong> FM1-43 was followed <strong>in</strong> real time<br />

dur<strong>in</strong>g second period <strong>of</strong> stimulation with high K + after dye<br />

load<strong>in</strong>g and wash<strong>in</strong>g. Upon stimulation, the NT2 neurons<br />

released the dye with<strong>in</strong> 5 m<strong>in</strong> (Fig. 3d–g) <strong>in</strong>dicat<strong>in</strong>g the presynaptic<br />

vesicle exocytosis. The unload<strong>in</strong>g <strong>of</strong> FM1-43 also<br />

depended on the presence <strong>of</strong> calcium <strong>in</strong> the stimulation buffer<br />

(Fig. 3g and h). To confirm that the FM1-43Fx loaded punctate<br />

sta<strong>in</strong><strong>in</strong>g <strong>in</strong>dicates pre-synaptic structure, we performed<br />

co-sta<strong>in</strong><strong>in</strong>g <strong>of</strong> FM1-43Fx loaded neurons with the synaptotagm<strong>in</strong><br />

I antibody (Fig. 4a–c). Even though the sta<strong>in</strong><strong>in</strong>g for<br />

FM1-43Fx weakened dur<strong>in</strong>g permeabilization and wash<strong>in</strong>g <strong>of</strong><br />

the culture, we found that most <strong>of</strong> the FM-labeled puncta were<br />

synaptotagm<strong>in</strong> I-immunoreactive (Fig. 4c).<br />

NO and cyclic GMP modulates synaptic vesicle recycl<strong>in</strong>g<br />

and exocytosis<br />

The expression <strong>of</strong> nNOS by human NT2 neurons was<br />

detected by western blott<strong>in</strong>g which resolved a major prote<strong>in</strong><br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446


(a)<br />

(d)<br />

(g) (h)<br />

Fig. 3 <strong>Human</strong> NT2 neurons display synaptic vesicle recycl<strong>in</strong>g and<br />

exocytosis. NT2 neurons were cultured for 28 DIV and labeled with the<br />

fixable analog <strong>of</strong> FM1-43 (FM1-43Fx) and normal FM1-43. (a, b) FM1-<br />

43Fx is loaded <strong>in</strong>to recycl<strong>in</strong>g synaptic vesicle <strong>of</strong> NT2 neurons upon<br />

stimulation with high K + compared to basal (KRH). Photomicrographs<br />

at the bottom <strong>of</strong> (a, b) represent the respective phase contrast images<br />

<strong>of</strong> NT2 neurons. (c) Quantification <strong>of</strong> FM1-43Fx uptake as fluorescent<br />

<strong>in</strong>tensity <strong>of</strong> the basal level <strong>in</strong>dicates significantly higher uptake <strong>of</strong> the<br />

dye <strong>in</strong>to NT2 neuron term<strong>in</strong>als upon stimulation by high K + as compared<br />

with KRH and Ca 2+ free high K + . (d–h) The unload<strong>in</strong>g <strong>of</strong> normal<br />

FM1-43 from NT2 neurons was followed upon stimulation with high K +<br />

band around 155 kDa (Fig. 5a). This is <strong>in</strong> l<strong>in</strong>e with previous<br />

reports (Lee et al. 2001; Tegenge and Bicker 2009) for the<br />

expression <strong>of</strong> nNOS <strong>in</strong> the NT2 precursor cells. Subpopulations<br />

<strong>of</strong> NT2 neurons positively sta<strong>in</strong>ed for the nNOS<br />

monoclonal antibody (Fig. 5b). The immun<strong>of</strong>luorescence<br />

label<strong>in</strong>g for nNOS appeared ma<strong>in</strong>ly around the cell somata<br />

and weakly along the neurites <strong>of</strong> NT2 neurons (Fig. 5b). To<br />

reveal the presence <strong>of</strong> the functional NO receptor sGC, we<br />

used an antibody aga<strong>in</strong>st cGMP (De Vente et al. 1987). In<br />

(b)<br />

(c)<br />

(e) (f)<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446<br />

NO and cyclic nucleotide modulates vesicle release | 1439<br />

with or without Ca 2+ . (d) FM1-43 loaded punctate sta<strong>in</strong><strong>in</strong>g after<br />

wash<strong>in</strong>g with KRH (10 m<strong>in</strong>) and calcium free buffer (5 m<strong>in</strong>). (e, f) FM1-<br />

43 loaded term<strong>in</strong>als undergo desta<strong>in</strong><strong>in</strong>g with<strong>in</strong> 5 m<strong>in</strong> upon stimulation<br />

with high K + . Arrow heads <strong>in</strong> (d–f) <strong>in</strong>dicates representative puncta that<br />

undergo desta<strong>in</strong><strong>in</strong>g. (g) A representative FM1-43 desta<strong>in</strong><strong>in</strong>g curve<br />

upon stimulation by high K + with or without calcium. (h) The percent<br />

fluorescence <strong>in</strong>tensity rema<strong>in</strong><strong>in</strong>g at the end <strong>of</strong> stimulation with high K +<br />

was significantly lower than Ca 2+ free high K + . Data represent<br />

mean ± SEM <strong>of</strong> at least 100 (c) and 20 (h) synaptic puncta from three<br />

<strong>in</strong>dependent experiments. Scale bars: 50 lm.<br />

the absence <strong>of</strong> an exogenous source <strong>of</strong> NO, on average<br />

6.7 ± 1% cGMP-IR cells were detected (Fig. 5c). The<br />

number <strong>of</strong> cGMP-IR cells <strong>in</strong>creased significantly up to<br />

49.4 ± 4% upon stimulation with the NO donor SNP<br />

(Fig. 5d). The number <strong>of</strong> NO-<strong>in</strong>duced cGMP-IR cells<br />

reduced to 23.6 ± 4% when SNP stimulation was accompanied<br />

by the sGC <strong>in</strong>hibitor, ODQ (Fig. 5e). Our data clearly<br />

demonstrate the expression <strong>of</strong> a NO-sensitive sGC <strong>in</strong><br />

subpopulations <strong>of</strong> NT2 neurons.


1440 | M. A. Tegenge et al.<br />

(a) (b) (c)<br />

Fig. 4 FM1-43Fx loaded synaptic puncta co-localize with synaptotagm<strong>in</strong> I. NT2 neurons (21 DIV) were loaded with (a) FM1-43Fx, fixed and<br />

sta<strong>in</strong>ed aga<strong>in</strong>st (b) cytoplasmic doma<strong>in</strong> <strong>of</strong> synaptotagm<strong>in</strong> I. (c) FM1-43Fx loaded synaptic puncta were colocalized with synaptotagm<strong>in</strong> I. Scale bar:<br />

25 lm.<br />

(a)<br />

(b)<br />

(c) (d) (e)<br />

Fig. 5 NT2 neurons express nNOS and functional sGC. (a) Western<br />

blott<strong>in</strong>g <strong>of</strong> cell lysate from NT2 precursor cells and NT2 neurons. The<br />

antibody aga<strong>in</strong>st nNOS recognizes a prote<strong>in</strong> band <strong>of</strong> apparent<br />

molecular weight at 155 kDa and the lower band represents acetylated<br />

a-tubul<strong>in</strong>. (b) Immun<strong>of</strong>luorescence detection <strong>of</strong> nNOS <strong>in</strong> cell somata <strong>of</strong><br />

NT2 neurons and around the neurites. (c–e) cGMP-IR <strong>of</strong> NT2 neurons<br />

for the detection <strong>of</strong> functional sGC. NT2 neurons cultured for 28 DIV<br />

was exposed for 20 m<strong>in</strong> to SNP (1 mM) or SNP + ODQ (50 lM) together<br />

with 1 mM 3-isobutyl-1-methylxanth<strong>in</strong>e (IBMX), as a phos-<br />

To <strong>in</strong>vestigate the <strong>in</strong>volvement <strong>of</strong> the NO-cGMP signal<br />

transduction pathway <strong>in</strong> pre-synaptic neurotransmitter release,<br />

we exposed NT2 neurons to the NO donor SNP. This<br />

resulted <strong>in</strong> a significantly higher <strong>in</strong>corporation <strong>of</strong> the lum<strong>in</strong>al<br />

phodiesterase <strong>in</strong>hibitor and 20 lM <strong>of</strong> 3-(50-Hydroxymethyl-20-furyl)-1benzyl<br />

<strong>in</strong>dazole (YC-1), as an enhancer <strong>of</strong> NO-<strong>in</strong>duced activity <strong>of</strong> sGC.<br />

Controls were also treated with IBMX and YC-1. Cultures were fixed<br />

with 4% paraformaldehyde <strong>in</strong> PBS and sta<strong>in</strong>ed aga<strong>in</strong>st cGMP. (c)<br />

Under control conditions, NT2 neurons showed little cGMP-IR. (d) The<br />

addition <strong>of</strong> the NO donor SNP <strong>in</strong>creased the cGMP-IR. (e) Application<br />

<strong>of</strong> SNP together with ODQ reduced NO-<strong>in</strong>duced cGMP-IR. Blue <strong>in</strong> (c–<br />

e) <strong>in</strong>dicates nuclear counter-sta<strong>in</strong><strong>in</strong>g (4¢,6-diamid<strong>in</strong>o-2¢-phenyl<strong>in</strong>doldihydrochloride).<br />

Scale bar: 50 lm (b), 100 lm (c–e).<br />

synaptotagm<strong>in</strong> I antibody similar to depolarization-<strong>in</strong>duced<br />

label<strong>in</strong>g by high K + (Fig. 6a). The <strong>in</strong>corporation <strong>of</strong> antilum<strong>in</strong>al<br />

synaptotagm<strong>in</strong> by SNP was blocked when it was<br />

applied together with the sGC <strong>in</strong>hibitor, ODQ (Fig. 6a).<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446


(a)<br />

(b)<br />

(c)<br />

Fig. 6 NO and cGMP analog facilitate synaptic vesicle recycl<strong>in</strong>g. (a)<br />

The NO donor, SNP (1 mM) significantly <strong>in</strong>creased the <strong>in</strong>corporation<br />

<strong>of</strong> lum<strong>in</strong>al synaptotagm<strong>in</strong> I <strong>in</strong>to NT2 neuron that was blocked by the<br />

sGC <strong>in</strong>hibitor, ODQ (50 lM). (b) The uptake <strong>of</strong> FM1-43Fx was significantly<br />

<strong>in</strong>creased by SNP (1 mM). Pre-<strong>in</strong>cubation with sGC <strong>in</strong>hibitor<br />

(ODQ, 50 lM) blocked SNP-<strong>in</strong>duced uptake <strong>of</strong> FM1-43Fx. (c) The cell<br />

membrane-permeable analog <strong>of</strong> cGMP, 8-bromoguanos<strong>in</strong>e-3¢, 5¢cyclic<br />

monophosphate, sodium salt enhanced the uptake <strong>of</strong> FM1-<br />

43Fx. Data represent mean ± SEM <strong>of</strong> percent fluorescence <strong>in</strong>tensity<br />

<strong>of</strong> at least 80 synaptic puncta from three <strong>in</strong>dependent experiments.<br />

Likewise, the uptake <strong>of</strong> FM1-43Fx was enhanced by SNP<br />

(Fig. 6b) and a membrane-permeable analog <strong>of</strong> cGMP<br />

(Fig. 6c). Pre-<strong>in</strong>cubation with ODQ resulted <strong>in</strong> block<strong>in</strong>g <strong>of</strong><br />

SNP-<strong>in</strong>duced uptake <strong>of</strong> the dye (Fig. 6b). These experimental<br />

results suggest that NO/cGMP modulates synaptic vesicle<br />

recycl<strong>in</strong>g <strong>in</strong> NT2 neurons.<br />

The unload<strong>in</strong>g <strong>of</strong> FM1-43 was followed <strong>in</strong> real time by<br />

monitor<strong>in</strong>g the dimm<strong>in</strong>g <strong>of</strong> fluorescence <strong>in</strong>tensity after<br />

stimulation with two NO donors, SNP, and NOC-12. Both<br />

NO donors <strong>in</strong>duced the release <strong>of</strong> the dye with a desta<strong>in</strong><strong>in</strong>g<br />

pr<strong>of</strong>ile <strong>of</strong> FM1-43 similar to that seen <strong>in</strong> high K + (Fig. 7a).<br />

At the end <strong>of</strong> 5 m<strong>in</strong> last<strong>in</strong>g stimulation with SNP or NOC-<br />

12, the percentage <strong>of</strong> fluorescence rema<strong>in</strong><strong>in</strong>g <strong>in</strong> NT2 nerve<br />

term<strong>in</strong>als was significantly lower than <strong>in</strong> normal KRH,<br />

<strong>in</strong>dicat<strong>in</strong>g that NO-<strong>in</strong>duced pre-synaptic vesicle exocytosis<br />

(Fig. 7b). After FM1-43 was loaded <strong>in</strong>to the neurons, the<br />

culture was <strong>in</strong>cubated with ODQ for 10 m<strong>in</strong> before the<br />

unload<strong>in</strong>g <strong>of</strong> FM1-43 was <strong>in</strong>itiated with SNP or NOC-12.<br />

Pre-<strong>in</strong>cubation with ODQ significantly blocked SNP<br />

(Fig. 7b) and NOC-12 (Fig. 7c) <strong>in</strong>duced FM1-43 unload<strong>in</strong>g.<br />

These results show that NO <strong>in</strong>duces vesicle exocytosis <strong>in</strong><br />

NT2 neurons via the cGMP pathway.<br />

cAMP/PKA signal transduction modulates synaptic vesicle<br />

recycl<strong>in</strong>g and exocytosis<br />

To reveal a possible <strong>in</strong>volvement <strong>of</strong> the cAMP signal<strong>in</strong>g<br />

pathway, we used activators and <strong>in</strong>hibitors <strong>of</strong> this cascade <strong>in</strong><br />

FM1-43 uptake experiments. The uptake <strong>of</strong> FM1-43Fx was<br />

significantly enhanced <strong>in</strong> the presence <strong>of</strong> the adenylyl<br />

cyclase stimulator forskol<strong>in</strong> (Fig. 8a) and a cell membrane<br />

permeable analog <strong>of</strong> cAMP (8-Br-cAMP) (Fig. 8b). Pre<strong>in</strong>cubation<br />

<strong>of</strong> NT2 neurons with PKA <strong>in</strong>hibitors (Rp-cAMP<br />

or H-89) significantly reduced forskol<strong>in</strong>-<strong>in</strong>duced (Fig. 8c)<br />

and high K + -<strong>in</strong>duced (data not shown) uptake <strong>of</strong> FM1-43Fx<br />

<strong>in</strong>dicat<strong>in</strong>g that cAMP/PKA pathway positively regulates<br />

synaptic vesicle recycl<strong>in</strong>g. Additionally, we followed the<br />

release <strong>of</strong> FM1-43 upon stimulation with forskol<strong>in</strong>. Stimulation<br />

<strong>of</strong> NT2 neurons with forskol<strong>in</strong> resulted <strong>in</strong> rapid<br />

release <strong>of</strong> FM1-43 which <strong>in</strong>dicates that forskol<strong>in</strong> <strong>in</strong>duced<br />

vesicle exocytosis (Fig. 8d). Pre-<strong>in</strong>cubation with H-89<br />

significantly reduced the release <strong>of</strong> FM1-43 <strong>in</strong>dicat<strong>in</strong>g the<br />

participation <strong>of</strong> PKA <strong>in</strong> forskol<strong>in</strong> <strong>in</strong>duced vesicle exocystosis<br />

(Fig. 8d).<br />

Discussion<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446<br />

NO and cyclic nucleotide modulates vesicle release | 1441<br />

<strong>Human</strong> NT2 neurons display synaptic vesicle recycl<strong>in</strong>g and<br />

pre-synaptic release<br />

Recently, we have shown that the human NT2 cells<br />

proliferate dur<strong>in</strong>g ret<strong>in</strong>oic acid treatment as spherical<br />

aggregate culture and cells migrate out <strong>of</strong> the aggregate<br />

to acquire fully differentiated neuronal phenotypes (Tegenge<br />

and Bicker 2009). In this study, we followed the progress <strong>of</strong><br />

NT2 neuronal maturation after the migratory phase when<br />

neuronal networks are formed. After 7 DIV, <strong>in</strong>tense MAP2<br />

sta<strong>in</strong><strong>in</strong>g <strong>in</strong>dicated full differentiation <strong>of</strong> NT2 neurons but<br />

sta<strong>in</strong><strong>in</strong>g for phosphorylated Tau, <strong>in</strong>dicative <strong>of</strong> axonogenesis,<br />

was weak (Fig. 1a and b). However, after 28 DIV,<br />

<strong>in</strong>tense Tau-sta<strong>in</strong><strong>in</strong>g appeared on long neurites, whereas<br />

MAP2 sta<strong>in</strong><strong>in</strong>g did not change <strong>in</strong> the cell cultures (Fig. 1e<br />

and f).


1442 | M. A. Tegenge et al.<br />

(a)<br />

(b)<br />

(c)<br />

Fig. 7 NO <strong>in</strong>duces pre-synaptic vesicle exocytosis via the cGMP<br />

pathway. (a) Representative desta<strong>in</strong><strong>in</strong>g pr<strong>of</strong>iles <strong>of</strong> FM1-43 after stimulation<br />

by the NO donors SNP (1 mM) and N-Ethyl-2-(1-ethyl-2-hydroxy-2-nitrosohydraz<strong>in</strong>o)<br />

ethanam<strong>in</strong>e (NOC-12) (100 lM) <strong>in</strong><br />

comparison with high K + . (b) The percent fluorescence <strong>in</strong>tensity<br />

rema<strong>in</strong><strong>in</strong>g after 5 m<strong>in</strong> stimulation with SNP was comparable with<br />

60 mM KCl. Pre-<strong>in</strong>cubation with ODQ (50 lM) for 10 m<strong>in</strong> and stimulation<br />

with SNP + ODQ for 5 m<strong>in</strong> resulted <strong>in</strong> block<strong>in</strong>g <strong>of</strong> SNP-<strong>in</strong>duced<br />

FM1-43 unload<strong>in</strong>g. (c) NOC-12 significantly <strong>in</strong>duced unload<strong>in</strong>g <strong>of</strong><br />

FM1-43 which was blocked by ODQ. Data represent mean ± SEM <strong>of</strong><br />

percent fluorescence <strong>in</strong>tensity rema<strong>in</strong><strong>in</strong>g after 5 m<strong>in</strong> stimulation from<br />

at least 15 synaptic puncta.<br />

Immunocytochemical evidence for pre-synaptic maturation<br />

<strong>of</strong> NT2 neurons was monitored by sta<strong>in</strong><strong>in</strong>g for the presynaptic<br />

prote<strong>in</strong>s, synaps<strong>in</strong> I, and synaptotagm<strong>in</strong> I. Synaps<strong>in</strong><br />

I is a member <strong>of</strong> the synaps<strong>in</strong> family specifically associated<br />

with synaptic vesicles which has been implicated <strong>in</strong> the<br />

synapse development and regulation <strong>of</strong> neurotransmitter<br />

release (reviewed by de Camilli et al. 1990; Hilfiker et al.<br />

1999; Ferreira and Rapoport 2002). Synaptotagm<strong>in</strong> I is the<br />

best characterized is<strong>of</strong>orm <strong>of</strong> the synaptotagm<strong>in</strong> family <strong>of</strong><br />

vesicle-associated prote<strong>in</strong>s, thought to function as a calcium<br />

sensor for fast neurotransmitter release at the synapse<br />

(reviewed by Yoshihara and Montana 2004; Chapman 2008).<br />

Immun<strong>of</strong>luorescence sta<strong>in</strong><strong>in</strong>g and quantification <strong>of</strong> synaptic<br />

puncta showed that the level <strong>of</strong> synaps<strong>in</strong> and synaptotagm<strong>in</strong><br />

I <strong>in</strong>creased significantly with the length <strong>of</strong> <strong>in</strong> vitro<br />

culture. Synaps<strong>in</strong> immunoreactivity became more <strong>in</strong>tense<br />

and appeared <strong>in</strong> the neuronal process with<strong>in</strong> 14–28 DIV<br />

(Fig. 1g). Previously, for NT2 neurons generated by the<br />

classical methods, <strong>in</strong>tense immunoreactivity to synaps<strong>in</strong> was<br />

observed upon co-culture with rat astrocytes (Hartley et al.<br />

1999). Transcriptional up-regulation <strong>of</strong> synaps<strong>in</strong>s dur<strong>in</strong>g<br />

ret<strong>in</strong>oic acid-<strong>in</strong>duced differentiation <strong>of</strong> NT2 cells has been<br />

reported (Leypoldt et al. 2002). The immunosta<strong>in</strong><strong>in</strong>g <strong>of</strong><br />

synaptotagm<strong>in</strong> I that appeared as punctate sta<strong>in</strong><strong>in</strong>g (Fig. 1h)<br />

could <strong>in</strong>dicate local accumulation <strong>of</strong> the prote<strong>in</strong> at presumptive<br />

pre-synaptic sites for participation <strong>in</strong> fast neurotransmitter<br />

release. The expression <strong>of</strong> other pre-synaptic prote<strong>in</strong>s<br />

such as synaptobrev<strong>in</strong>, synaptophys<strong>in</strong>, and SNAP25 has also<br />

been reported for NT2 neurons (Sheridan and Maltese 1998).<br />

However, it rema<strong>in</strong>s unclear whether these pre-synaptic<br />

prote<strong>in</strong>s express<strong>in</strong>g immunopositive puncta represent functional<br />

synaptic term<strong>in</strong>als. In this study, we showed for the<br />

first time that human NT2 neurons display synaptic vesicle<br />

recycl<strong>in</strong>g and exocytosis by two <strong>in</strong>dependent approaches.<br />

Firstly, we used functional immun<strong>of</strong>luorescence methods to<br />

label pre-synaptic term<strong>in</strong>als <strong>of</strong> NT2 neurons dur<strong>in</strong>g synaptic<br />

vesicle recycl<strong>in</strong>g. For this purpose, an antibody directed to<br />

the lum<strong>in</strong>al doma<strong>in</strong> <strong>of</strong> synaptotagm<strong>in</strong> I was used <strong>in</strong> the<br />

presence <strong>of</strong> high K + . It is well known that high K + <strong>in</strong>duces<br />

multiple rounds <strong>of</strong> exocytosis lead<strong>in</strong>g to label<strong>in</strong>g <strong>of</strong> the entire<br />

pool <strong>of</strong> recycl<strong>in</strong>g synaptic vesicles (Kl<strong>in</strong>gauf et al. 1998;<br />

Sara et al. 2002; Menegon et al. 2006). In NT2 neurons, the<br />

label<strong>in</strong>g <strong>of</strong> synaptic vesicles by anti-lum<strong>in</strong>al synaptotagm<strong>in</strong> I<br />

depends on depolarization <strong>in</strong>duced by high K + and presence<br />

<strong>of</strong> calcium <strong>in</strong> the stimulation buffer (Fig. 2c). Our data also<br />

show that the extent <strong>of</strong> synaptic vesicle recycl<strong>in</strong>g depends on<br />

the length <strong>of</strong> <strong>in</strong> vitro culture correspond<strong>in</strong>g to the expression<br />

<strong>of</strong> the pre-synaptic prote<strong>in</strong>s (Figs. 1i, j and 2b).<br />

In the second approach, synaptic vesicle recycl<strong>in</strong>g and<br />

exocytosis was analyzed by employ<strong>in</strong>g the fluorescent dye,<br />

FM1-43. The protocol for label<strong>in</strong>g pre-synaptic vesicles that<br />

display synaptic vesicle recycl<strong>in</strong>g by FM1-43 is well<br />

established (Cochilla et al. 1999; Gaffield and Betz 2006).<br />

The uptake <strong>of</strong> FM1-43Fx was significantly <strong>in</strong>creased upon<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446


Fig. 8 The cAMP/PKA pathway modulates<br />

synaptic vesicle recycl<strong>in</strong>g <strong>in</strong> NT2 neurons.<br />

The uptake <strong>of</strong> FM1-43Fx was significantly<br />

<strong>in</strong>creased <strong>in</strong> the presence <strong>of</strong> (a) forskol<strong>in</strong><br />

and (b) cAMP analog (8-Br-cAMP). (c) Pre<strong>in</strong>cubation<br />

<strong>of</strong> NT2 neurons for 30 m<strong>in</strong> with<br />

prote<strong>in</strong> k<strong>in</strong>ase A antagonist (10 lM<br />

RPcAMP or H-89) significantly reduced<br />

forskol<strong>in</strong> <strong>in</strong>duced FM1-43Fx uptake. (d)<br />

Real time imag<strong>in</strong>g <strong>of</strong> vesicle exocytosis<br />

upon stimulation with forskol<strong>in</strong> (100 lM)<br />

<strong>in</strong>dicates fast release <strong>of</strong> FM1-43 and the<br />

percent fluorescence <strong>in</strong>tensity rema<strong>in</strong><strong>in</strong>g<br />

after 5 m<strong>in</strong> was comparable with that <strong>of</strong><br />

high K + . Pre-<strong>in</strong>cubation with H-89 (10 lM)<br />

for 10 m<strong>in</strong> and stimulation with forskol<strong>in</strong><br />

(100 lM) resulted <strong>in</strong> marked reduction <strong>of</strong><br />

the release <strong>of</strong> FM1-43. Data represent<br />

mean ± SEM at least 80 (a–c) and 10 (d)<br />

synaptic puncta.<br />

stimulation with high K + <strong>in</strong> the presence <strong>of</strong> calcium. The<br />

little punctate sta<strong>in</strong><strong>in</strong>g that appeared upon stimulation with<br />

KRH (basal) dur<strong>in</strong>g the <strong>in</strong>corporation <strong>of</strong> lum<strong>in</strong>al synaptotagm<strong>in</strong><br />

I antibody (Fig. 2c) and FM1-43Fx (Fig. 3a) could<br />

<strong>in</strong>dicate spontaneous activity <strong>of</strong> the neurons. Intrigu<strong>in</strong>gly,<br />

confocal images <strong>of</strong> NT2 neurons loaded with FM1-43Fx and<br />

double-labeled for synaptotagm<strong>in</strong> I display clear co-localization<br />

<strong>of</strong> the punctate sta<strong>in</strong><strong>in</strong>g (Fig. 4a–c). Thus, both<br />

methods <strong>in</strong>dicate that human NT2 neurons undergo synaptic<br />

vesicle recycl<strong>in</strong>g <strong>in</strong> a depolarization and calcium-dependent<br />

manner.<br />

Unload<strong>in</strong>g <strong>of</strong> FM1-43 which monitors pre-synaptic vesicle<br />

exocytosis revealed rapid release <strong>of</strong> the dye <strong>in</strong> a both<br />

depolarization and calcium-dependent manner (Fig. 3d–h).<br />

This pre-synaptic vesicle exocytosis <strong>in</strong>duced by high K +<br />

presumably <strong>in</strong>dicates neurotransmitter release at nerve<br />

term<strong>in</strong>als. Recently, we have shown that NT2 neurons<br />

express markers for glutamate, GABA, and for chol<strong>in</strong>ergic<br />

transmission (Podrygajlo et al. 2009) suggest<strong>in</strong>g that these<br />

neurotransmitters are putatively released by high K + . At the<br />

post-synaptic site, expression <strong>of</strong> functional receptors <strong>in</strong>clud<strong>in</strong>g<br />

GABA A-receptors (Matsuoka et al. 1997; Neelands et al.<br />

1998, 1999;.), NMDA-type (Rootwelt et al. 1998; Paquet-<br />

Durand and Bicker 2004; Paquet-Durand et al. 2006; Garcia<br />

de Arriba et al. 2006), and a-am<strong>in</strong>o-3-hydroxy-5-methylisoxazole-4-propionate<br />

(AMPA)-type glutamate receptors<br />

(Rootwelt et al. 1998) have been reported. Nevertheless,<br />

functional synapses between NT2 neurons have been demonstrated<br />

electrophysiologically only when cultured together<br />

(a) (b)<br />

(c) (d)<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446<br />

NO and cyclic nucleotide modulates vesicle release | 1443<br />

with rat astrocytes (Hartley et al. 1999). Spontaneous but<br />

uncorrelated fir<strong>in</strong>g patterns <strong>of</strong> NT2 neurons have been<br />

recorded on microelectrode arrays (Görtz et al. 2004). Here,<br />

we showed pre-synaptic vesicle release <strong>in</strong> NT2 neurons <strong>in</strong><br />

the absence <strong>of</strong> glia cells. S<strong>in</strong>ce we could demonstrate that the<br />

neurons display <strong>in</strong>creas<strong>in</strong>g levels <strong>of</strong> an axonal marker (Tau),<br />

pre-synaptic prote<strong>in</strong>s (synaps<strong>in</strong> and synaptotagm<strong>in</strong> I), and<br />

synaptic vesicle recycl<strong>in</strong>g with time <strong>in</strong> culture, NT2 neurons<br />

seem to undergo pre-synaptic maturation processes. Therefore,<br />

NT2 neurons can be utilized as a model to <strong>in</strong>vestigate<br />

the cellular and molecular basis <strong>of</strong> synaptic vesicle recycl<strong>in</strong>g<br />

and pre-synaptic vesicle release <strong>in</strong> human nerve cells.<br />

NO signal transduction and pre-synaptic vesicle release<br />

<strong>Nitric</strong> oxide synthase has been shown to be expressed dur<strong>in</strong>g<br />

neural development and synaptogenesis (Williams et al.<br />

1994; Ogilvie et al. 1995; Sporns and Jenk<strong>in</strong>son 1997; Gibbs<br />

and Truman 1998) <strong>in</strong>dicat<strong>in</strong>g that NO is <strong>in</strong>volved <strong>in</strong> synaptic<br />

maturation processes. This study demonstrates for the first<br />

time that NO and cyclic nucleotides regulate vesicle<br />

recycl<strong>in</strong>g and pre-synaptic vesicle release <strong>in</strong> human model<br />

neurons. Subpopulations <strong>of</strong> NT2 neurons positively sta<strong>in</strong>ed<br />

for the nNOS monoclonal antibody (Fig. 5b). Thus, similar<br />

to other neurotransmitter phenotypes (Guillema<strong>in</strong> et al.<br />

2000; Podrygajlo et al. 2009), the NT2 neurons are also<br />

heterogeneous with respect to nNOS expression. These<br />

neurons express<strong>in</strong>g nNOS could serve as endogenous source<br />

<strong>of</strong> NO. The presence <strong>of</strong> functional sGC was demonstrated by<br />

exogenous stimulation <strong>of</strong> the neurons with NO donor and


1444 | M. A. Tegenge et al.<br />

subsequent detection <strong>of</strong> cGMP by immun<strong>of</strong>luorescence<br />

methods (Fig. 5c–e). NO <strong>in</strong>duced cGMP-IR was reduced <strong>in</strong><br />

the presence <strong>of</strong> a sGC <strong>in</strong>hibitor.<br />

The NO donor, SNP significantly <strong>in</strong>creased the label<strong>in</strong>g <strong>of</strong><br />

NT2 neurons by lum<strong>in</strong>al synaptotagm<strong>in</strong> I which was<br />

blocked by the sGC <strong>in</strong>hibitor ODQ <strong>in</strong>dicat<strong>in</strong>g that NO<br />

facilitates synaptic vesicle recycl<strong>in</strong>g via the cGMP pathway.<br />

SNP enhanced the uptake <strong>of</strong> FM1-43Fx similar to stimulation<br />

with high K + . The effect <strong>of</strong> exogenous NO application<br />

was blocked by ODQ. Moreover, a membrane permeable<br />

analog <strong>of</strong> cGMP significantly <strong>in</strong>creased the uptake <strong>of</strong> FM1-<br />

43Fx which further confirms the participation <strong>of</strong> cGMP<br />

pathway <strong>in</strong> NO-mediated synaptic vesicle recycl<strong>in</strong>g. These<br />

experimental results suggest that the NO/cGMP signal<strong>in</strong>g<br />

pathway could facilitate synaptic vesicle recycl<strong>in</strong>g <strong>in</strong> human<br />

neurons. A recent study demonstrates that cGMP reduces the<br />

cycle time for synaptic vesicles through the enhancement <strong>of</strong><br />

vesicular traffic rate from the recycl<strong>in</strong>g pool to the readily<br />

releasable pool and accelerates fast endocytosis (Petrov<br />

et al. 2008). Pre-synaptic exocytosis was followed <strong>in</strong> real<br />

time dur<strong>in</strong>g stimulation <strong>of</strong> NT2 neurons by two NO donors,<br />

SNP and NOC-12. Both SNP and NOC-12 <strong>in</strong>duced the<br />

unload<strong>in</strong>g <strong>of</strong> FM1-43 similar to high K + (Fig. 7a–c). The<br />

unload<strong>in</strong>g <strong>of</strong> FM1-43 by SNP and NOC-12 was blocked by<br />

ODQ which <strong>in</strong>dicates that NO <strong>in</strong>duces vesicle release via the<br />

cGMP pathway. Our data are <strong>in</strong> agreement with several<br />

reports that demonstrated <strong>in</strong>creased pre-synaptic transmitter<br />

release by the NO/cGMP pathway (Arancio et al. 1995,<br />

2001; Lu et al. 1999; Wildemann and Bicker 1999; Li et al.<br />

2004; Nickels et al. 2007). To analyze whether the NO/<br />

cGMP <strong>in</strong>duced enhancement <strong>of</strong> vesicle exocytosis is caused<br />

by excitatory effects at the level <strong>of</strong> the membrane or by<br />

directly facilitat<strong>in</strong>g the calcium-dependent release mach<strong>in</strong>ery,<br />

electrophysiological techniques will be required. Moreover,<br />

<strong>in</strong> hippocampal neurons and synaptosomes, NO has<br />

been shown to <strong>in</strong>duce neurotransmitter release <strong>in</strong>dependent<br />

<strong>of</strong> <strong>in</strong>tracellular calcium levels (Meffert et al. 1994; Sporns<br />

and Jenk<strong>in</strong>son 1997).<br />

One potential <strong>in</strong>tracellular downstream effector prote<strong>in</strong> for<br />

the cGMP signal<strong>in</strong>g pathway is PKG. Enhanced transmitter<br />

release via phosphorylation <strong>of</strong> pre-synaptic voltage-gated K +<br />

channels by PKG has recently been implicated (Yang et al.<br />

2007). Additionally, formation <strong>of</strong> new cluster <strong>of</strong> pre- and<br />

post-synaptic prote<strong>in</strong>s which <strong>in</strong>volves phosporylation <strong>of</strong> PKG<br />

substrate prote<strong>in</strong>s, vasodilator-stimulated phosphoprote<strong>in</strong>,<br />

and RhoA have been reported (Wang et al. 2005). A direct<br />

action <strong>of</strong> NO without the <strong>in</strong>volvement <strong>of</strong> cGMP and PKG via<br />

S-nitrosylation on the exo–endocytotic mach<strong>in</strong>eries has been<br />

also implicated (Meffert et al. 1996; Ahern et al. 2002). Even<br />

though our data from both immun<strong>of</strong>luorescence and FM<br />

imag<strong>in</strong>g strongly support the participation <strong>of</strong> the cGMP<br />

pathway, we could not completely exclude the direct action <strong>of</strong><br />

NO s<strong>in</strong>ce the sGC <strong>in</strong>hibitor, ODQ did only partially block the<br />

unload<strong>in</strong>g <strong>of</strong> FM1-43 <strong>in</strong>duced by NOC-12 (Fig. 7c).<br />

Pre-synaptic activation <strong>of</strong> calcium-sensitive adenylyl<br />

cyclase that leads to a rise <strong>in</strong> the level <strong>of</strong> cAMP and<br />

consequent activation <strong>of</strong> PKA has been implicated to<br />

enhance the probability <strong>of</strong> neurotransmitter release (Siegelbaum<br />

et al. 1982; Chavez-Noriega and Stevens 1994;<br />

Trudeau et al. 1996; Evans and Morgan 2003). The prote<strong>in</strong>s<br />

<strong>in</strong>volved <strong>in</strong> pre-synaptic vesicle exocytosis have been<br />

suggested as major substrates for PKA-dependent phosphorylation<br />

(reviewed by Leenders and Sheng 2005). Phosphorylation<br />

<strong>of</strong> synaps<strong>in</strong> by the cAMP/PKA signal<strong>in</strong>g which leads<br />

to release <strong>of</strong> vesicles from the reserve pool has been<br />

suggested to enhance pre-synaptic vesicle releases (Fiumara<br />

et al. 2004, 2007; Bonanomi et al. 2005; Menegon et al.<br />

2006).<br />

Here, we showed that stimulation <strong>of</strong> adenylyl cyclase by<br />

forskol<strong>in</strong> <strong>in</strong>duces vesicle exocytosis and facilitates vesicle<br />

recycl<strong>in</strong>g while <strong>in</strong>hibition <strong>of</strong> PKA by Rp-cAMP or H-89<br />

blocks the effect <strong>of</strong> forskol<strong>in</strong> <strong>in</strong>dicat<strong>in</strong>g that the cAMP/PKA<br />

pathway modulates vesicle release and synaptic vesicle<br />

recycl<strong>in</strong>g <strong>in</strong> NT2 neurons. Our result is <strong>in</strong> l<strong>in</strong>e with the<br />

<strong>in</strong>creased neurotransmitter release reported for hippocampal<br />

neurons upon activation <strong>of</strong> cAMP/PKA signal<strong>in</strong>g (Trudeau<br />

et al. 1996, 1998). However, we do not know at present<br />

which target prote<strong>in</strong>s are modified by NO and cyclic<br />

nucleotide signal<strong>in</strong>g <strong>in</strong> NT2 neurons. In our readily accessible<br />

human model neurons, we can now <strong>in</strong>vestigate the<br />

candidate prote<strong>in</strong>s modulated via NO and cyclic nucleotide<br />

signal transduction.<br />

In summary, we have presented a model <strong>of</strong> human neurons<br />

that express <strong>in</strong>creas<strong>in</strong>g levels <strong>of</strong> pre-synaptic prote<strong>in</strong>s and<br />

display synaptic vesicle recycl<strong>in</strong>g. Our data also revealed for<br />

the first time that the NO and cyclic nucleotide signal<br />

transduction modulates synaptic vesicle recycl<strong>in</strong>g and exocytosis<br />

<strong>in</strong> human model neurons.<br />

Acknowledgements<br />

We thank Dr J. de Vente for his k<strong>in</strong>d gift <strong>of</strong> the cGMP antiserum, Dr<br />

S. Knipp for the discussion dur<strong>in</strong>g the preparation <strong>of</strong> the manuscript<br />

and S. Tan for technical support. This research was <strong>in</strong> part supported<br />

by BMBF grant 0313732 to G. Bicker and by a DFG grant (FG<br />

1103, BI 262/16-1). M. A. Tegenge received a Georg-Christoph-<br />

Lichtenberg scholarship from the M<strong>in</strong>istry for Science and Culture<br />

<strong>of</strong> Lower Saxony.<br />

References<br />

Ahern G. P., Klyachko V. A. and Jackson M. B. (2002) cGMP and Snitrosylation:<br />

two routes for modulation <strong>of</strong> neuronal excitability by<br />

NO. Trends Neurosci. 25, 510–517.<br />

Andrews P. W. (1984) Ret<strong>in</strong>oic acid <strong>in</strong>duces neuronal differentiation <strong>of</strong> a<br />

cloned human embryonal carc<strong>in</strong>oma cell l<strong>in</strong>e <strong>in</strong> vitro. Dev. Biol.<br />

103, 285–293.<br />

Arancio O., Kandel E. R. and Hawk<strong>in</strong>s R. D. (1995) Activity-dependent<br />

long-term enhancement <strong>of</strong> transmitter release by presynaptic 3¢, 5¢cyclic<br />

GMP <strong>in</strong> cultured hippocampal neurons. Nature 376, 74–80.<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446


Arancio O., Kiebler M., Lee C. J., Lev-Ram V., Tsien R. Y., Kandel E. R.<br />

and Hawk<strong>in</strong>s R. D. (1996) <strong>Nitric</strong> oxide acts directly <strong>in</strong> the presynaptic<br />

neuron to produce long-term potentiation <strong>in</strong> cultured<br />

hippocampal neurons. Cell 87, 1025–1035.<br />

Arancio O., Antonova I., Gambaryan S., Lohmann S. M., Wood J. S.,<br />

Lawrence D. S. and Hawk<strong>in</strong>s R. D. (2001) Presynaptic role <strong>of</strong><br />

cGMP-dependent prote<strong>in</strong> k<strong>in</strong>ase dur<strong>in</strong>g long-last<strong>in</strong>g potentiation.<br />

J. Neurosci. 21, 143–149.<br />

Bicker G. (2005) STOP and GO with NO: nitric oxide as a regulator <strong>of</strong><br />

cell motility <strong>in</strong> simple bra<strong>in</strong>s. BioEssays 27, 495–505.<br />

Bicker G. (2007) Pharmacological approaches to nitric oxide signall<strong>in</strong>g<br />

dur<strong>in</strong>g neural development <strong>of</strong> locusts and other model <strong>in</strong>sects.<br />

Arch. Insect Biochem. Physiol. 64, 43–58.<br />

Boehn<strong>in</strong>g D. and Snyder S. H. (2003) Novel neural modulators. Annu.<br />

Rev. Neurosci. 26, 105–131.<br />

Bonanomi D., Menegon A., Miccio A., Ferrari G., Corradi A., Kao H.,<br />

Benfenati F. and Valtorta F. (2005) Phosphorylation <strong>of</strong> synaps<strong>in</strong> I<br />

by cAMP-dependent prote<strong>in</strong> k<strong>in</strong>ase controls synaptic vesicle<br />

dynamics <strong>in</strong> develop<strong>in</strong>g neurons. J. Neurosci. 25, 7299–7308.<br />

Camilli P. D., Benfenati F., Valtorta F. and Greengard P. (1990) The<br />

synaps<strong>in</strong>s. Annu. Rev. Cell Biol. 6, 433–460.<br />

Chapman E. R. (2008) How does synaptotagm<strong>in</strong> trigger neurotransmitter<br />

release? Annu. Rev. Biochem. 77, 615–641.<br />

Chavez-Noriega L. E. and Stevens C. F. (1994) Increased transmitter<br />

release at excitatory synapses produced by direct activation <strong>of</strong><br />

adenylate cyclase <strong>in</strong> rat hippocampal slices. J. Neurosci. 14, 310–<br />

317.<br />

Cochilla A. J., Angleson J. K. and Betz W. J. (1999) Monitor<strong>in</strong>g<br />

secretory membrane with FM1-43 fluorescence. Annu. Rev.<br />

Neurosci. 22, 1–10.<br />

De Vente J. D., Ste<strong>in</strong>busch H. W. and Schipper J. (1987) A new approach<br />

to immunocytochemistry <strong>of</strong> 3¢, 5¢-cyclic guanos<strong>in</strong>e monophosphate:<br />

preparation, specificity, and <strong>in</strong>itial application <strong>of</strong> a new<br />

antiserum aga<strong>in</strong>st formaldehyde-fixed 3¢, 5¢-cyclic guanos<strong>in</strong>e<br />

monophosphate. Neuroscience 22, 361–373.<br />

Enikolopov G., Banerji J. and Kuz<strong>in</strong> B. (1999) <strong>Nitric</strong> oxide and Drosophila<br />

development. Cell Death Differ. 6, 956–963.<br />

Evans G. J. O. and Morgan A. (2003) Regulation <strong>of</strong> the exocytotic<br />

mach<strong>in</strong>ery by cAMP-dependent prote<strong>in</strong> k<strong>in</strong>ase: implications for<br />

presynaptic plasticity. Biochem. Soc. Trans. 31, 824–827.<br />

Ferreira A. and Rapoport M. (2002) The synaps<strong>in</strong>s: beyond the regulation<br />

<strong>of</strong> neurotransmitter release. Cell. Mol. Life Sci. 59, 589–595.<br />

Fiumara F., Giovedì S., Menegon A., Milanese C., Merlo D., Montarolo<br />

P. G., Valtorta F., Benfenati F. and Ghirardi M. (2004) Phosphorylation<br />

by cAMP-dependent prote<strong>in</strong> k<strong>in</strong>ase is essential for synaps<strong>in</strong>-<strong>in</strong>duced<br />

enhancement <strong>of</strong> neurotransmitter release <strong>in</strong><br />

<strong>in</strong>vertebrate neurons. J. Cell Sci. 117, 5145–5154.<br />

Fiumara F., Milanese C., Corradi A., Giovedì S., Leit<strong>in</strong>ger G., Menegon<br />

A., Montarolo P. G., Benfenati F. and Ghirardi M. (2007) Phosphorylation<br />

<strong>of</strong> synaps<strong>in</strong> doma<strong>in</strong> A is required for post-tetanic<br />

potentiation. J. Cell Sci. 120, 3228–3237.<br />

Gaffield M. A. and Betz W. J. (2006) Imag<strong>in</strong>g synaptic vesicle exocytosis<br />

and endocytosis with FM dyes. Nat. Protoc. 1, 2916–2921.<br />

Gao Z. Y., Xu G., Stwora-Wojczyk M. M., Matsch<strong>in</strong>sky F. M., Lee V. M.<br />

and Wolf B. A. (1998) Ret<strong>in</strong>oic acid <strong>in</strong>duction <strong>of</strong> calcium channel<br />

expression <strong>in</strong> human NT2N neurons. Biochem. Biophys. Res.<br />

Commun. 247, 407–413.<br />

Garcia de Arriba S., Wegner F., Grüner K., Verdaguer E., Pallas M.,<br />

Cam<strong>in</strong>s A., Wagner A., Wohlfahrt K. and Allgaier C. (2006) Different<br />

capacities <strong>of</strong> various NMDA receptor antagonists to prevent<br />

ischemia-<strong>in</strong>duced neurodegeneration <strong>in</strong> human cultured NT2 neurons.<br />

Neurochem. Int. 49, 466–474.<br />

Garthwaite J. (2008) Concepts <strong>of</strong> neural nitric oxide-mediated transmission.<br />

Eur. J. Neurosci. 27, 2783–2802.<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446<br />

NO and cyclic nucleotide modulates vesicle release | 1445<br />

Gibbs S. M. and Truman J. W. (1998) <strong>Nitric</strong> oxide and cyclic GMP<br />

regulate ret<strong>in</strong>al pattern<strong>in</strong>g <strong>in</strong> the optic lobe <strong>of</strong> Drosophila. Neuron<br />

20, 83–93.<br />

Görtz P., Fleischer W., Rosenbaum C., Otto F. and Siebler M. (2004)<br />

Neuronal network properties <strong>of</strong> human teratocarc<strong>in</strong>oma cell l<strong>in</strong>ederived<br />

neurons. Bra<strong>in</strong> Res. 1018, 18–25.<br />

Guillema<strong>in</strong> I., Alonso G., Patey G., Privat A. and Chaudieu I. (2000)<br />

<strong>Human</strong> NT2 neurons express a large variety <strong>of</strong> neurotransmission<br />

phenotypes <strong>in</strong> vitro. J. Comp. Neurol. 422, 380–395.<br />

Hartley R. S., Margulis M., Fishman P. S., Lee V. M. and Tang C. M.<br />

(1999) Functional synapses are formed between human NTera2<br />

(NT2N, hNT) neurons grown on astrocytes. J. Comp. Neurol. 407,<br />

1–10.<br />

Hawk<strong>in</strong>s R. D., Zhuo M. and Arancio O. (1994) <strong>Nitric</strong> oxide and carbon<br />

monoxide as possible retrograde messengers <strong>in</strong> hippocampal longterm<br />

potentiation. J. Neurobiol. 25, 652–665.<br />

Hilfiker S., Pieribone V. A., Czernik A. J., Kao H. T., August<strong>in</strong>e G. J.<br />

and Greengard P. (1999) Synaps<strong>in</strong>s as regulators <strong>of</strong> neurotransmitter<br />

release. Philos. Trans. R. Soc. Lond. B Biol. Sci. 354, 269–<br />

279.<br />

Kl<strong>in</strong>gauf J., Kavalali E. T. and Tsien R. W. (1998) K<strong>in</strong>etics and regulation<br />

<strong>of</strong> fast endocytosis at hippocampal synapses. Nature 394,<br />

581–585.<br />

Kraszewski K., Mundigl O., Daniell L., Verderio C., Matteoli M. and<br />

Camilli P. D. (1995) Synaptic vesicle dynamics <strong>in</strong> liv<strong>in</strong>g cultured<br />

hippocampal neurons visualized with CY3-conjugated antibodies<br />

directed aga<strong>in</strong>st the lumenal doma<strong>in</strong> <strong>of</strong> synaptotagm<strong>in</strong>. J. Neurosci.<br />

15, 4328–4342.<br />

Lee M., Hyun D. H., Jenner P. and Halliwell B. (2001) Effect <strong>of</strong><br />

proteasome <strong>in</strong>hibition on cellular oxidative damage, antioxidant<br />

defences and nitric oxide production. J. Neurochem. 78, 32–41.<br />

Leenders A. G. M. and Sheng Z. (2005) Modulation <strong>of</strong> neurotransmitter<br />

release by the second messenger-activated prote<strong>in</strong> k<strong>in</strong>ases:<br />

implications for presynaptic plasticity. Pharmacol. Ther. 105,<br />

69–84.<br />

Leypoldt F., Flajolet M. and Methner A. (2002) Neuronal differentiation<br />

<strong>of</strong> cultured human NTERA-2cl.D1 cells leads to <strong>in</strong>creased<br />

expression <strong>of</strong> synaps<strong>in</strong>s. Neurosci. Lett. 324, 37–40.<br />

Li D., Chen S. and Pan H. (2002) <strong>Nitric</strong> oxide <strong>in</strong>hibits sp<strong>in</strong>ally project<strong>in</strong>g<br />

paraventricular neurons through potentiation <strong>of</strong> presynaptic<br />

GABA release. J. Neurophysiol. 88, 2664–2674.<br />

Li D., Chen S., F<strong>in</strong>negan T. F. and Pan H. (2004) <strong>Signal</strong>l<strong>in</strong>g pathway <strong>of</strong><br />

nitric oxide <strong>in</strong> synaptic GABA release <strong>in</strong> the rat paraventricular<br />

nucleus. J. Physiol. 554, 100–110.<br />

Lu Y. F., Kandel E. R. and Hawk<strong>in</strong>s R. D. (1999) <strong>Nitric</strong> oxide signal<strong>in</strong>g<br />

contributes to late-phase LTP and CREB phosphorylation <strong>in</strong> the<br />

hippocampus. J. Neurosci. 19, 10250–10261.<br />

Matsuoka T., Kondoh T., Tamaki N. and Nishizaki T. (1997) The GABAA<br />

receptor is expressed <strong>in</strong> human neurons derived from a teratocarc<strong>in</strong>oma<br />

cell l<strong>in</strong>e. Biochem. Biophys. Res. Commun. 237, 719–723.<br />

Matteoli M., Takei K., Per<strong>in</strong> M. S., Südh<strong>of</strong> T. C. and De Camilli P.<br />

(1992) Exo-endocytotic recycl<strong>in</strong>g <strong>of</strong> synaptic vesicles <strong>in</strong> develop<strong>in</strong>g<br />

processes <strong>of</strong> cultured hippocampal neurons. J. Cell Biol. 117,<br />

849–861.<br />

Meffert M. K., Premack B. A. and Schulman H. (1994) <strong>Nitric</strong> oxide<br />

stimulates Ca (2+)-<strong>in</strong>dependent synaptic vesicle release. Neuron<br />

12, 1235–1244.<br />

Meffert M. K., Calakos N. C., Scheller R. H. and Schulman H. (1996)<br />

<strong>Nitric</strong> oxide modulates synaptic vesicle dock<strong>in</strong>g fusion reactions.<br />

Neuron 16, 1229–1236.<br />

Menegon A., Bonanomi D., Albert<strong>in</strong>azzi C., Lotti F., Ferrari G., Kao H.,<br />

Benfenati F., Baldelli P. and Valtorta F. (2006) Prote<strong>in</strong> k<strong>in</strong>ase Amediated<br />

synaps<strong>in</strong> I phosphorylation is a central modulator <strong>of</strong><br />

Ca2+-dependent synaptic activity. J. Neurosci. 26, 11670–11681.


1446 | M. A. Tegenge et al.<br />

Micheva K. D., Buchanan J., Holz R. W. and Smith S. J. (2003) Retrograde<br />

regulation <strong>of</strong> synaptic vesicle endocytosis and recycl<strong>in</strong>g.<br />

Nat. Neurosci. 6, 905–906.<br />

Neelands T. R., Greenfield L. J., Zhang J., Turner R. S. and Macdonald<br />

R. L. (1998) GABAA receptor pharmacology and subtype mRNA<br />

expression <strong>in</strong> human neuronal NT2-N cells. J. Neurosci. 18, 4993–<br />

5007.<br />

Neelands T. R., Zhang J. and Macdonald R. L. (1999) GABA (A)<br />

receptors expressed <strong>in</strong> undifferentiated human teratocarc<strong>in</strong>oma<br />

NT2 cells differ from those expressed by differentiated NT2-N<br />

cells. J. Neurosci. 19, 7057–7065.<br />

Neelands T. R., K<strong>in</strong>g A. P. and Macdonald R. L. (2000) Functional<br />

expression <strong>of</strong> L-, N-, P/Q-and R-type calcium channels <strong>in</strong> the<br />

human NT2-N cell l<strong>in</strong>e. J. Neurophysiol. 84, 2933–2944.<br />

Nickels T. J., Reed G. W., Drummond J. T., Blev<strong>in</strong>s D. E., Lutz M. C.<br />

and Wilson D. F. (2007) Does nitric oxide modulate transmitter<br />

release at the mammalian neuromuscular junction? Cl<strong>in</strong>. Exp.<br />

Pharmacol. Physiol. 34, 318–326.<br />

Ogilvie P., Schill<strong>in</strong>g K., Bill<strong>in</strong>gsley M. L. and Schmidt H. H. (1995)<br />

Induction and variants <strong>of</strong> neuronal nitric oxide synthase type I<br />

dur<strong>in</strong>g synaptogenesis. FASEB J. 9, 799–806.<br />

Paquet-Durand F. and Bicker G. (2004) Hypoxic/ischaemic cell damage<br />

<strong>in</strong> cultured human NT-2 neurons. Bra<strong>in</strong> Res. 1011, 33–47.<br />

Paquet-Durand F., Gierse A. and Bicker G. (2006) Diltiazem protects<br />

human NT-2 neurons aga<strong>in</strong>st excitotoxic damage <strong>in</strong> a model <strong>of</strong><br />

simulated ischemia. Bra<strong>in</strong> Res. 1124, 45–54.<br />

Paquet-Durand F., Tan S. and Bicker G. (2003) Turn<strong>in</strong>g teratocarc<strong>in</strong>oma<br />

cells <strong>in</strong>to neurons: rapid differentiation <strong>of</strong> NT-2 cells <strong>in</strong> float<strong>in</strong>g<br />

spheres. Bra<strong>in</strong> Res. Dev. Bra<strong>in</strong> Res. 142, 161–167.<br />

Petrov A. M., G<strong>in</strong>iatull<strong>in</strong> A. R., Sitdikova G. F. and Zefirov A. L. (2008)<br />

The role <strong>of</strong> cGMP-dependent signal<strong>in</strong>g pathway <strong>in</strong> synaptic vesicle<br />

cycle at the frog motor nerve term<strong>in</strong>als. J. Neurosci. 28, 3216–<br />

3222.<br />

Pleasure S. J., Page C. and Lee V. M. (1992) Pure, postmitotic, polarized<br />

human neurons derived from NTera 2 cells provide a system for<br />

express<strong>in</strong>g exogenous prote<strong>in</strong>s <strong>in</strong> term<strong>in</strong>ally differentiated neurons.<br />

J. Neurosci. 12, 1802–1815.<br />

Podrygajlo G., Tegenge M. A., Gierse A., Paquet-Durand F., Tan S.,<br />

Bicker G. and Stern M. (2009) Cellular phenotypes <strong>of</strong> human<br />

model neurons (NT2) after differentiation <strong>in</strong> aggregate culture. Cell<br />

Tissue Res. 336, 439–452.<br />

Rootwelt T., Dunn M., Yudk<strong>of</strong>f M., Itoh T., Almaas R. and Pleasure D.<br />

(1998) Hypoxic cell death <strong>in</strong> human NT2-N neurons: <strong>in</strong>volvement<br />

<strong>of</strong> NMDA and non-NMDA glutamate receptors. J. Neurochem. 71,<br />

1544–1553.<br />

Sara Y., Mozhayeva M. G., Liu X. and Kavalali E. T. (2002) Fast vesicle<br />

recycl<strong>in</strong>g supports neurotransmission dur<strong>in</strong>g susta<strong>in</strong>ed stimulation<br />

at hippocampal synapses. J. Neurosci. 22, 1608–1617.<br />

Sheridan K. M. and Maltese W. A. (1998) Expression <strong>of</strong> Rab3A GTPase<br />

and other synaptic prote<strong>in</strong>s is <strong>in</strong>duced <strong>in</strong> differentiated NT2N<br />

neurons. J. Mol. Neurosci. 10, 121–128.<br />

Siegelbaum S. A., Camardo J. S. and Kandel E. R. (1982) Seroton<strong>in</strong> and<br />

cyclic AMP close s<strong>in</strong>gle K + channels <strong>in</strong> Aplysia sensory neurones.<br />

Nature 299, 413–417.<br />

Sporns O. and Jenk<strong>in</strong>son S. (1997) Potassium ion- and nitric oxide<strong>in</strong>duced<br />

exocytosis from populations <strong>of</strong> hippocampal synapses<br />

dur<strong>in</strong>g synaptic maturation <strong>in</strong> vitro. Neuroscience 80, 1057–1073.<br />

Taqatqeh F., Mergia E., Neitz A., Eysel U. T., Koesl<strong>in</strong>g D. and Mittmann<br />

T. (2009) More than a retrograde messenger: nitric oxide needs two<br />

cGMP pathways to <strong>in</strong>duce hippocampal long-term potentiation.<br />

J. Neurosci. 29, 9344–9350.<br />

Tegenge M. A. and Bicker G. (2009) <strong>Nitric</strong> oxide and cGMP signal<br />

transduction positively regulates the motility <strong>of</strong> human neuronal<br />

precursor (NT2) cells. J. Neurochem. 110, 1828–1841.<br />

Trudeau L. E., Emery D. G. and Haydon P. G. (1996) Direct modulation<br />

<strong>of</strong> the secretory mach<strong>in</strong>ery underlies PKA-dependent synaptic<br />

facilitation <strong>in</strong> hippocampal neurons. Neuron 17, 789–797.<br />

Trudeau L. E., Fang Y. and Haydon P. G. (1998) Modulation <strong>of</strong> an early<br />

step <strong>in</strong> the secretory mach<strong>in</strong>ery <strong>in</strong> hippocampal nerve term<strong>in</strong>als.<br />

Proc. Natl Acad. Sci. USA 95, 7163–7168.<br />

Verderio C., Grumelli C., Raiteri L., Coco S., Paluzzi S., Cacc<strong>in</strong> P.,<br />

Rossetto O., Bonanno G., Montecucco C. and Matteoli M. (2007)<br />

Traffic <strong>of</strong> botul<strong>in</strong>um tox<strong>in</strong>s A and E <strong>in</strong> excitatory and <strong>in</strong>hibitory<br />

neurons. Traffic 8, 142–153.<br />

Wang H., Lu F., J<strong>in</strong> I., Udo H., Kandel E. R., Vente J. D., Walter U.,<br />

Lohmann S. M., Hawk<strong>in</strong>s R. D. and Antonova I. (2005) Presynaptic<br />

and postsynaptic roles <strong>of</strong> NO, cGK, and RhoA <strong>in</strong> long-last<strong>in</strong>g potentiation<br />

and aggregation <strong>of</strong> synaptic prote<strong>in</strong>s. Neuron 45, 389–403.<br />

Wildemann B. and Bicker G. (1999) <strong>Nitric</strong> oxide and cyclic GMP <strong>in</strong>duce<br />

vesicle release at Drosophila neuromuscular junction. J. Neurobiol.<br />

39, 337–346.<br />

Williams C. V., Nordquist D. and McLoon S. C. (1994) Correlation <strong>of</strong> nitric<br />

oxide synthase expression with chang<strong>in</strong>g patterns <strong>of</strong> axonal projections<br />

<strong>in</strong> the develop<strong>in</strong>g visual system. J. Neurosci. 14, 1746–1755.<br />

Yang Q., Chen S. R., Li D. P. and Pan H. L. (2007) Kv1.1/1.2 channels<br />

are downstream effectors <strong>of</strong> nitric oxide on synaptic GABA release<br />

to preautonomic neurons <strong>in</strong> the paraventricular nucleus. Neuroscience<br />

149, 315–327.<br />

Yoshihara M. and Montana E. S. (2004) The synaptotagm<strong>in</strong>s: calcium<br />

sensors for vesicular traffick<strong>in</strong>g. Neuroscientist 10, 566–574.<br />

Ó 2009 The Authors<br />

Journal Compilation Ó 2009 International Society for Neurochemistry, J. Neurochem. (2009) 111, 1434–1446


<strong>Nitric</strong> oxide signal<strong>in</strong>g as regulator <strong>of</strong> human neuronal progenitor cell migration<br />

Million Adane Tegenge 1,2 , Thomas D<strong>in</strong>o Rockel 3 , Ellen Fritsche* 3,4 , Gerd Bicker* 1,2<br />

1. Division <strong>of</strong> Cell Biology, Institute <strong>of</strong> Physiology, University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e Hannover,<br />

Bisch<strong>of</strong>sholer Damm 15, D-30173 Hannover, Germany<br />

2. Center for Systems Neuroscience (ZSN), Hannover, Germany<br />

3. Group <strong>of</strong> Toxicology, Institut für Umweltmediz<strong>in</strong>ische Forschung gGmbH at the He<strong>in</strong>rich He<strong>in</strong>e-<br />

University, Düsseldorf, Germany<br />

4. Cl<strong>in</strong>ic <strong>of</strong> Dermatology, RWTH Aachen, Germany<br />

*Correspond<strong>in</strong>g authors: Gerd Bicker, Division <strong>of</strong> Cell Biology, Institute <strong>of</strong> Physiology, University<br />

<strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e Hannover, Bisch<strong>of</strong>sholer Damm 15, D-30173 Hannover, Germany, Tel.:<br />

+49 5 11 856 7765, Fax: +49 5 11 856 7687<br />

e-mail: gerd.bicker@tiho-hannover.de<br />

Ellen Fritsche, Group <strong>of</strong> Toxicology, Institut für Umweltmediz<strong>in</strong>ische Forschung gGmbH at the<br />

He<strong>in</strong>rich He<strong>in</strong>e-University, Auf’m Hennekamp 50, 40225, Düsseldorf, Germany, Tel. +49<br />

2113389217, Fax: +49 211 3190910<br />

(In preparation)<br />

e-mail: ellen.fritsche@uni-duesseldorf.de


Abstract<br />

Neuronal migration is a cellular event that is critical for the development <strong>of</strong> functional neuronal<br />

network. A number <strong>of</strong> genes responsible for human congential disorders featur<strong>in</strong>g neuronal<br />

migration defects have been identified. In this study we exam<strong>in</strong>e the role <strong>of</strong> nitric oxide (NO)<br />

signal<strong>in</strong>g <strong>in</strong> the migration <strong>of</strong> fetal human neural progenitor cells (hNPCs) which was cultured as<br />

neurospheres. Cells migrate out <strong>of</strong> the human neurospheres and differentiate <strong>in</strong>to both neuronal and<br />

glial cells. The migrat<strong>in</strong>g cells express functional guanylyl cyclase (sGC) that produces cGMP upon<br />

activation with NO, an effect reduced <strong>in</strong> the presence <strong>of</strong> sGC <strong>in</strong>hibitor. Application <strong>of</strong> enzyme<br />

<strong>in</strong>hibitors <strong>of</strong> sGC and prote<strong>in</strong> k<strong>in</strong>ase G (PKG) blocked the migration <strong>of</strong> cells out <strong>of</strong> human<br />

neurospheres. Inhibition <strong>of</strong> sGC can be rescued by a membrane permeable analogue <strong>of</strong> cGMP. In<br />

ga<strong>in</strong> <strong>of</strong> function experiments both NO donor and cGMP analogue facilitate cell migration<br />

suggest<strong>in</strong>g that NO-cGMP signal<strong>in</strong>g positively regulates hNPCs migration. These results provide<br />

first experimental evidence for a role <strong>of</strong> NO-cGMP signal transduction as a regulator <strong>of</strong> cell<br />

migration dur<strong>in</strong>g early development <strong>of</strong> the human bra<strong>in</strong>.<br />

Key words: bra<strong>in</strong> development, human neurospheres, NO, cGMP<br />

2


Introduction<br />

Neuronal precursor cell migration dur<strong>in</strong>g early phase <strong>of</strong> development and <strong>in</strong> some area <strong>of</strong> the adult<br />

bra<strong>in</strong> is important for the formation <strong>of</strong> functional networks <strong>in</strong> the bra<strong>in</strong>. The gaseous messenger<br />

nitric oxide (NO), which is produced by nitric oxide synthase enzymes (NOS), appears to play<br />

critical role <strong>in</strong> proliferation, migration and synaptogenesis dur<strong>in</strong>g nerve cell development<br />

(Enikolopov et al., 1999; Cárdenas et al., 2005; Bicker, 2005). A major receptor for NO is the<br />

cGMP synthesis<strong>in</strong>g enzyme soluble guanylyl cyclase (sGC) (Garthwaite, 2008). NO signal<br />

transduction has been shown to regulate the migration <strong>of</strong> cerebellar neurons, <strong>in</strong>sect enteric neurons,<br />

and early develop<strong>in</strong>g Xenopus neuronal cells (Tanaka et al., 1994; Haase and Bicker, 2003;<br />

Peunova et al., 2007; Knipp and Bicker, 2009). Neuroanatomical studies us<strong>in</strong>g markers aga<strong>in</strong>st<br />

NOS and sGC <strong>in</strong>dentify the migrat<strong>in</strong>g neuroblasts <strong>of</strong> the rostral migratory stream as potential<br />

targets for NO signal<strong>in</strong>g <strong>in</strong> the adult bra<strong>in</strong> (Moreno-López et al., 2000; Gutierrez-Mec<strong>in</strong>as et al.,<br />

2007). In a model <strong>of</strong> ischemic stroke, NO donor has been shown to facilitate neuroblasts migration<br />

<strong>in</strong> subventricular zone and dentate gyrus (Zhang et al., 2001; Cui et al., 2009).<br />

The transient expression <strong>of</strong> NOS dur<strong>in</strong>g development has provided additional evidence for an<br />

<strong>in</strong>volvement <strong>of</strong> NO signal<strong>in</strong>g <strong>in</strong> neuronal migration. In the develop<strong>in</strong>g rat bra<strong>in</strong>, nNOS has been<br />

demonstrated to be expressed transiently, reach<strong>in</strong>g the highest level between embryonic day 16 and<br />

postnatal day 0. These periods corresponds to the migration <strong>of</strong> neuronal precursors from the<br />

ventricular zone to the external layers <strong>of</strong> the cortex (Bredt and Snyder, 1994; Nott et al., 2008). In<br />

the fetal human sp<strong>in</strong>al cord, some neurons express NOS as they migrate to their f<strong>in</strong>al dest<strong>in</strong>ation<br />

(Foster and Phelps, 2000). Recently, we have used cultured human NT2 cells to demonstrate that<br />

NO and cGMP signal transduction regulate the migration <strong>of</strong> neuronal precursors (Tegenge and<br />

Bicker, 2009). However, there is no functional evidence <strong>in</strong>dicat<strong>in</strong>g the participation <strong>of</strong> NO signal<strong>in</strong>g<br />

dur<strong>in</strong>g early development <strong>of</strong> human nervous system.<br />

The expansion <strong>of</strong> multi-potent human neural progenitor cells <strong>in</strong> vitro may <strong>of</strong>fer a model system to<br />

study the molecular and cellular basis <strong>of</strong> human bra<strong>in</strong> development. In this study we used human<br />

neural progenitor cells (hNPCs) ma<strong>in</strong>ta<strong>in</strong>ed as three-dimensional free-float<strong>in</strong>g neurospheres, to<br />

<strong>in</strong>vestigate the role <strong>of</strong> NO/cGMP signal transduction <strong>in</strong> neuronal migration. We used<br />

immunocytochemistry to show that hNPCs express neuronal NOS (nNOS) and functional sGC. By<br />

employ<strong>in</strong>g specific enzyme activators and <strong>in</strong>hibitors we provide the first experimental evidence that<br />

NO-cGMP signal transduction regulates cell migration <strong>in</strong> develop<strong>in</strong>g human bra<strong>in</strong> cells.<br />

Material and Methods<br />

NOC-18 [2, 2-(Hydroxynitrosohydraz<strong>in</strong>o) bis-ethanam<strong>in</strong>e] was purchased from Calbiochem<br />

(Darmstadt, Germany), 8-Br-cGMP (8-Bromoguanos<strong>in</strong>e 3’, 5’-cyclophosphate) and Rp-8-Br-cGMP<br />

3


(8-Bromoguanos<strong>in</strong>e-3’, 5’-cyclophosphothioate, Rp-isomer) were purchased from Alexis<br />

Biochemicals (Lörrach, Germany). All other substances were obta<strong>in</strong>ed from Sigma (Taufkirchen,<br />

Germany) unless otherwise noted.<br />

Cell culture<br />

The hNPCs used <strong>in</strong> this study were purchased from Lonza Verviers SPRL (Verviers, Belgium),<br />

which was obta<strong>in</strong>ed from three <strong>in</strong>dividuals. The hNPCs was cultured as previously described<br />

(Fritsche et al., 2005). Briefly, cells were cultured <strong>in</strong> proliferation medium (Dulbecco´s modified<br />

Eagle medium and Hams F12 (3:1) supplemented with B27 (Invitrogen GmBH, Karlsruhe,<br />

Germany), 20 ng/ml epidermal growth factor (EGF) (Biosource, Karlsruhe, Germany), 20 ng/ml<br />

recomb<strong>in</strong>ant human fibroblast growth factor (rhFGF) (R&D Systems, Wiesbaden-Nordenstadt,<br />

Germany), 100 U/ml penicill<strong>in</strong>, and 100 μg/ml streptomyc<strong>in</strong> <strong>in</strong> a humidified 92.5% air/7.5% CO2<br />

<strong>in</strong>cubator at 37°C <strong>in</strong> suspension culture. Medium was changed every 2–3 days and spheres were<br />

chopped with a McIlwa<strong>in</strong>e tissue chopper. Differentiation was <strong>in</strong>duced by growth factor withdrawal<br />

<strong>in</strong> the presence <strong>of</strong> DFN medium (Dulbecco modified Eagle medium and Hams F12 (3:1)<br />

supplemented with N2 (Invitrogen)) upon cultur<strong>in</strong>g the neurospheres on chamber slides (Becton<br />

Dick<strong>in</strong>son, Bedford, MA) coated with poly-D-lys<strong>in</strong>e and lam<strong>in</strong><strong>in</strong>.<br />

Migration assay<br />

hNPCs were seeded on poly-D-lys<strong>in</strong>e and lam<strong>in</strong><strong>in</strong> coated chamber slides at density <strong>of</strong> five<br />

neurospheres. The culture was treated with chemicals diluted <strong>in</strong> the differentiation medium for at<br />

least 72 h with a medium change after 48 h. The NO donor NOC-18, which decays with long half<br />

life (about 57 h at 22 o C) was prepared as 100 mM stock solution <strong>in</strong> 10 mM NaOH solution. ODQ<br />

(1H-[1, 2, 4]-oxadiazolo[4,3-a]qu<strong>in</strong>oxal<strong>in</strong>-1-one) was prepared as 20 mM stock solution <strong>in</strong> DMSO<br />

(dimethylsulphoxide). 8-Br-cGMP and Rp-8-Br-cGMP were directly dissolved <strong>in</strong> the medium. To<br />

determ<strong>in</strong>e the migration <strong>of</strong> cells out <strong>of</strong> neurospheres, images were acquired after 24 h <strong>of</strong> chemical<br />

application and migration distance was measured at four dist<strong>in</strong>ct locations from the rim <strong>of</strong> the<br />

spheres to furthest migrated cells (Moors et al., 2007; Tegenge and Bicker, 2009). The data were<br />

presented as mean ± s.e.m <strong>of</strong> three <strong>in</strong>dependent experiments.<br />

LDH cytotoxicity assay<br />

The lactate dehydrogenase (LDH) assay (CytoTox-One, Promega, Mannheim, Germany) was<br />

employed to assess cell death by measur<strong>in</strong>g the amount <strong>of</strong> LDH which is released <strong>in</strong>to the medium<br />

from dead cells. The assay was performed accord<strong>in</strong>g to the manufacturer’s <strong>in</strong>structions. Briefly,<br />

follow<strong>in</strong>g exposure to the chemicals for 48 h, supernatants <strong>of</strong> the culture medium (100 µl) was<br />

4


mixed with 50 µl <strong>of</strong> Cyto-Tox-One reagent and kept for 4 h <strong>in</strong> 96- well plate. Fluorescence <strong>in</strong>tensity<br />

was detected at excitation/emission wavelength <strong>of</strong> 540/590 nm. Results were presented as mean ±<br />

s.e.m <strong>of</strong> percent maximum LDH released from four measurements.<br />

Immunocytochemistry<br />

Immunocytochemical detection <strong>of</strong> βIII-tubul<strong>in</strong> (Sigma 1:10,000), nest<strong>in</strong> (Calbiochem 1:400) and<br />

glial fibrillary acidic prote<strong>in</strong> (GFAP, Sigma 1:500) was performed as previously described<br />

(Tegenge and Bicker, 2009). For the detection <strong>of</strong> cGMP immunoreactive (cGMP-IR) cells, hNPCs<br />

were pre<strong>in</strong>cubated for 20 m<strong>in</strong> at 37 o C with 1 mM sodium nitroprusside (SNP) as a NO donor, 20<br />

µM YC-1 (3-(50-Hydroxymethyl-20-furyl)-1-benzyl <strong>in</strong>dazole) as an enhancer <strong>of</strong> NO-<strong>in</strong>duced<br />

activity <strong>of</strong> sGC, 50 µM <strong>of</strong> ODQ as sGC <strong>in</strong>hibitor and 1 mM IBMX (3-isobutyl-1-methylxanth<strong>in</strong>e)<br />

as phosphodiesterase <strong>in</strong>hibitor. Cultures were washed once with PBS and fixed with 4% PFA. The<br />

polyclonal sheep cGMP antiserum (1:10,000; a k<strong>in</strong>d gift from Dr. J. de Vente, Maastricht<br />

University, Netherlands) was used as primary antibody to detect the level <strong>of</strong> cGMP.<br />

Microscopy and Data analysis<br />

Preparations were exam<strong>in</strong>ed with a fluorescent microscope Zeiss Axiovert 25 equipped with an<br />

Axiocam 3900 digital camera or Zeiss Axio Observer D1 (Zeiss, Göttengen, Germany). The<br />

percentage <strong>of</strong> nest<strong>in</strong>, βIII-tubul<strong>in</strong>, GFAP and cGMP-IR cell bodies was determ<strong>in</strong>ed divid<strong>in</strong>g the<br />

number <strong>of</strong> positively sta<strong>in</strong>ed cells by the total number <strong>of</strong> cells <strong>in</strong> the migration area. The data were<br />

presented as mean ± s.e.m <strong>of</strong> at least two <strong>in</strong>dependent experiments. Statistical comparisons <strong>of</strong><br />

control versus treatment were performed with the unpaired two-tailed Student’s t-test. Levels <strong>of</strong><br />

significance are <strong>in</strong>dicated as *P < 0.05, **P < 0.01 and ***P < 0.001.<br />

Results<br />

<strong>Nitric</strong> oxide-sensitive sGC is expressed <strong>in</strong> early develop<strong>in</strong>g human bra<strong>in</strong> cells<br />

On western blots a monoclonal antibody aga<strong>in</strong>st the nNOS labelled a major prote<strong>in</strong> band around<br />

155 KDa (Fig. 1), <strong>in</strong>dicat<strong>in</strong>g that nitric oxide could be endogenously produced early dur<strong>in</strong>g the<br />

development <strong>of</strong> human bra<strong>in</strong>. To reveal functional sGC that synthesizes cGMP upon activation by<br />

NO, we used an antibody aga<strong>in</strong>st cGMP (de Vente et al., 1987). In the absence <strong>of</strong> exogenous NO,<br />

only few cGMP-IR cells were detected (Fig. 2 A & D). The number <strong>of</strong> cGMP-IR cells <strong>in</strong>creased<br />

significantly up to 38.9 ± 2% upon stimulation with the NO donor SNP (Fig. 2 B & D). The NO<br />

<strong>in</strong>duced cGMP-IR cells were significantly reduced when SNP stimulation was accompanied by the<br />

sGC <strong>in</strong>hibitor ODQ (Fig. 2 C & D).<br />

5


Fig. 1 The antibody aga<strong>in</strong>st nNOS recognizes a prote<strong>in</strong> band <strong>of</strong> apparent molecular weight at 155<br />

kDa from the proliferat<strong>in</strong>g hNPCs on Western blot. NT2 spherical aggregates treated with ret<strong>in</strong>oic<br />

acid for 7 days was used as a positive control. The lower band represents acetylated-α-tubul<strong>in</strong>.<br />

Fig. 2 Functional NO/cGMP signal<strong>in</strong>g is present <strong>in</strong> early develop<strong>in</strong>g human nerve cells. The<br />

presence <strong>of</strong> NO-sensitive functional sGC was demonstrated by immunocytochemical detection <strong>of</strong><br />

cGMP. hNPCs were exposed for 20 m<strong>in</strong> to (A) 1 mM IBMX + 20 µM YC-1 (control), (B) SNP (1<br />

mM) or (C) SNP (1mM) + ODQ (50 µM) together with IBMX + YC-1. Cultures were fixed with<br />

4% PFA and sta<strong>in</strong>ed aga<strong>in</strong>st cGMP. (D) The cGMP-IR cells were counted after stimulation with<br />

NO donor (SNP) and compared with the control. (E) Schematic draw<strong>in</strong>g <strong>of</strong> NO/cGMP signal<br />

6


transduction together with an array <strong>of</strong> activators and <strong>in</strong>hibitors <strong>of</strong> the pathway. Scale bars: 50 µm.<br />

Data are mean ± s.e.m <strong>of</strong> 10 neurospheres from at least two <strong>in</strong>dependent experiments.<br />

cGMP-IR nest<strong>in</strong> and GFAP positive cells direct the migration <strong>of</strong> βIII-tubul<strong>in</strong>-IR cells<br />

The progress <strong>of</strong> neuronal differentiation with<strong>in</strong> hNPCs was monitored by immunocytochemical<br />

sta<strong>in</strong><strong>in</strong>g aga<strong>in</strong>st cytoskeletal markers <strong>of</strong> progenitor cells (nest<strong>in</strong>), early neurons (βIII-tubul<strong>in</strong>) and<br />

glial cells (GFAP). After 72 h <strong>of</strong> migration, the majority <strong>of</strong> the cells that migrate out <strong>of</strong> the<br />

neurospheres sta<strong>in</strong>ed for nest<strong>in</strong> with a few <strong>in</strong>terspersed neurons (Fig. 3A). A major subset <strong>of</strong> the<br />

cells was positive for the glial marker GFAP (Fig. 3A). Co-sta<strong>in</strong><strong>in</strong>g <strong>of</strong> cGMP with cytoskeletal<br />

markers <strong>in</strong>dicated that about 53.5 ± 2 % was GFAP-IR while the nest<strong>in</strong> and βIII-tubul<strong>in</strong>-IR cells<br />

constitute about 21 ± 3 and 9.6 ± 1 %, respectively (Fig. 3 B-D). Cells at the forefront <strong>of</strong> migration<br />

were ma<strong>in</strong>ly composed <strong>of</strong> the nest<strong>in</strong> and GFAP positive which are double-labeled for cGMP (Fig. 3<br />

E- G).<br />

cGMP/PKG signal<strong>in</strong>g positively regulates the migration <strong>of</strong> hNPCs<br />

After show<strong>in</strong>g the presence <strong>of</strong> a functional NO signal transduction pathway <strong>in</strong> the hNPCs we tested<br />

whether cGMP/PKG signal<strong>in</strong>g modulates the migration <strong>of</strong> cells out <strong>of</strong> the human neurospheres. The<br />

concentration <strong>of</strong> both <strong>in</strong>hibitors and activators <strong>of</strong> the pathway (Fig. 2 E) were selected based on<br />

previous experiments us<strong>in</strong>g migrat<strong>in</strong>g NT2 cells (Tegenge and Bicker, 2009). The sGC <strong>in</strong>hibitor<br />

(ODQ) significantly reduced the migration <strong>of</strong> cells out <strong>of</strong> neurospheres (Fig. 4 A, B). In the<br />

presence <strong>of</strong> PKG-I <strong>in</strong>hibitor (Rp-8-Br-cGMP), the migration <strong>of</strong> cell was reduced <strong>in</strong> a dose<br />

dependent manner (Fig. 4 C). Application <strong>of</strong> a cell membrane permeable analogue <strong>of</strong> cGMP, 8-BrcGMP<br />

partially rescued the block<strong>in</strong>g <strong>of</strong> cell migration by ODQ (Fig. 4 A, D). LDH cytotoxicity<br />

assay was conducted after the application <strong>of</strong> enzyme <strong>in</strong>hibitors <strong>of</strong> sGC and PKG. Both ODQ and<br />

Rp-8-Br-cGMP did not alter the viability <strong>of</strong> cells at the concentration used to <strong>in</strong>hibit cell migration<br />

(Fig. 4 E, F).<br />

7


Fig. 3 Immunocytochemical characterization <strong>of</strong> hNPCs for cytoskeletal markers and NO signal<br />

transduction. (A) After 72 h <strong>of</strong> migration on lam<strong>in</strong><strong>in</strong> coated chamber slides cells that migrate out <strong>of</strong><br />

human neurospheres express ma<strong>in</strong>ly nest<strong>in</strong> and GFAP with few <strong>in</strong>terspersed βIII-tubul<strong>in</strong>-IR cells.<br />

After exposure <strong>of</strong> hNPCs to SNP + IBMX + YC-1 for 20 m<strong>in</strong> culture was co-sta<strong>in</strong>ed for cGMP and<br />

(B) nest<strong>in</strong>, (C) βIII-tubul<strong>in</strong> or (D) GFAP. (E-G) cGMP-IR cells at the forefront <strong>of</strong> migration were<br />

co-sta<strong>in</strong>ed for nest<strong>in</strong> or GFAP but not for βIII-tubul<strong>in</strong>. Arrow heads (B-G) show representative colocalization<br />

<strong>of</strong> cGMP with respective cytoskeletal markers. Dense cellular aggregate sta<strong>in</strong><strong>in</strong>g (A-D)<br />

del<strong>in</strong>eates the edge <strong>of</strong> the neurospheres. Scale bar: 50 µm.<br />

NO and cGMP facilitates the migration <strong>of</strong> hNPCs<br />

Exposure <strong>of</strong> hNPCs to the NO donor NOC-18 (1-10 µM) for 24 h significantly facilitated the<br />

migration <strong>of</strong> cells out <strong>of</strong> the spheres (Fig. 5 A, B). However, 100 µM <strong>of</strong> NOC-18 significantly<br />

<strong>in</strong>hibited cell migration (Fig. 5 A, B). Parallel to cell migration experiments, we conducted cell<br />

viability assay upon NOC-18 application, which showed no sign <strong>of</strong> cytotoxicity (Fig. 5 C). A<br />

significant reduction <strong>of</strong> NO-<strong>in</strong>duced facilitation <strong>of</strong> cell migration was observed when NOC-18 was<br />

used <strong>in</strong> comb<strong>in</strong>ation with ODQ or Rp-8-Br-cGMP (Fig. 5 D). Application cGMP analogue, 8-BrcGMP<br />

significantly facilitated the migration <strong>of</strong> cells (Fig. 5 E).<br />

F<strong>in</strong>ally, we counted the number <strong>of</strong> βIII-tubul<strong>in</strong>-IR and GFAP-IR cells to exam<strong>in</strong>e the effects <strong>of</strong><br />

bioactive chemicals that activate and <strong>in</strong>hibit the sGC/PKG pathway. At the concentration used to<br />

8


modulate cell migration, none <strong>of</strong> the chemicals significantly altered the proportion <strong>of</strong> neuronal and<br />

glial marker (Fig. 6).<br />

Fig. 4 The sGC/PKG enzyme <strong>in</strong>hibitors block the migration <strong>of</strong> hNPCs. (A) Representative<br />

photomicrograph <strong>of</strong> hNPCs after 24 h <strong>of</strong> migration under control condition (0.25% DMSO) and <strong>in</strong><br />

the presence <strong>of</strong> ODQ or 8-Br-cGMP + ODQ. (B-D) Migration distance from the rim <strong>of</strong> the sphere<br />

to the furthest migrated cells was quantified at four dist<strong>in</strong>ct locations per sphere. Application <strong>of</strong> (B)<br />

ODQ and (C) Rp-8-Br-cGMP significantly blocked the migration hNPCs. (D) The blocked <strong>of</strong> cell<br />

migration by ODQ (50 µM) was rescued by 8-Br-cGMP (1 mM). Both (E) ODQ and (F) Rp-8-BrcGMP<br />

did not cause cytotoxicity. Data represent mean ± s.e.m <strong>of</strong> at least 10 neurospheres (B-D)<br />

and four measurements (E & F). Scale bar: 500 µm<br />

9


Fig. 5 NO donor and cell membrane permeable analogue <strong>of</strong> cGMP facilitates cell migration. (A)<br />

Representative photomicrographs <strong>of</strong> hNPCs after 24 h <strong>of</strong> migration under control condition and <strong>in</strong><br />

the presence <strong>of</strong> 10 µM NOC-18 and 100 µM NOC- 18. (B) The migration <strong>of</strong> cells <strong>in</strong>creased<br />

significantly <strong>in</strong> the presence <strong>of</strong> 1-10 µM <strong>of</strong> NOC-18. (C) The viability <strong>of</strong> cells was not altered by<br />

NOC-18. (D) Co-application <strong>of</strong> NOC-18 (10 µM) together with down-stream enzyme <strong>in</strong>hibitor <strong>of</strong><br />

sGC (ODQ, 50 µM) or PKG (Rp-8-Br-cGMP,1 µM) reduced NO-<strong>in</strong>duced migration <strong>of</strong> cells. (E) 8-<br />

10


Br-cGMP facilitated cell migration. Data represent mean ± s.e.m <strong>of</strong> 15 neurospheres (B, D, & E)<br />

and four measurements (C). Scale bar: 500 µm<br />

Fig. 6 The proportion <strong>of</strong> neuronal and glial cells was not changed <strong>in</strong> the presence <strong>of</strong> chemicals that<br />

modulate cell migration. Application ODQ (50 µM), Rp-8-Br-cGMP (1 µM), NOC-18 (10 µM) and<br />

8-Br-cGMP (1 mM) for 72 h <strong>in</strong> hNPCs culture did not significantly alter the percentage �III-tubul<strong>in</strong><br />

and GFAP-IR cells. Data are mean ± s.e.m. <strong>of</strong> 10 neurospheres from at least two <strong>in</strong>dependent<br />

experiments.<br />

Discussion<br />

hNPCs as a model for early develop<strong>in</strong>g human bra<strong>in</strong><br />

When the proliferat<strong>in</strong>g hNPCs were plated on lam<strong>in</strong><strong>in</strong> coated substrates, cells migrate out <strong>of</strong> the<br />

neurospheres. The migration distance from the rim <strong>of</strong> the spheres to the furthest migrated cells<br />

<strong>in</strong>creases with time <strong>in</strong> vitro and accompanied by differentiation <strong>in</strong>to neuronal and glial cells<br />

(Fritsche et al., 2005; Moors et al., 2007; 2009). Here, we performed immunocytochemical sta<strong>in</strong><strong>in</strong>g<br />

aga<strong>in</strong>st cytoskeletal markers <strong>of</strong> progenitor cells, early neurons and glial cells. The nest<strong>in</strong> and GFAP<br />

positive cells constitute the major proportion <strong>of</strong> cells that migrate out <strong>of</strong> the neurosphere with<strong>in</strong> 72 h<br />

(Fig. 2 A). Nevertheless, upon prolonged time <strong>of</strong> cultur<strong>in</strong>g (7-14 days) most <strong>of</strong> the nest<strong>in</strong> positive<br />

progenitor cells have been demonstrated to differentiate <strong>in</strong>to βIII-tubul<strong>in</strong>-IR cells (Brannen and<br />

Sugaya, 2000; Moors et al., 2009). S<strong>in</strong>ce hNPCs generate both neuronal and glial cells with the later<br />

<strong>in</strong> a higher proportion and at the forefront <strong>of</strong> migration (Fig. 3 G), the neurons appear to migrate<br />

along a glial scaffold. Indeed, real-time microscopic observation for 24 h revealed that cells migrate<br />

out <strong>of</strong> human neurosphere ma<strong>in</strong>ly by radial migration (Moors et al., 2009). Thus, the use <strong>of</strong> hNPCs<br />

to study cell migration may mimic aspects <strong>of</strong> radial neuroblast migration which accounts for 80-90<br />

% <strong>of</strong> cell migration dur<strong>in</strong>g early development <strong>of</strong> bra<strong>in</strong> (Ayala et al., 2007).<br />

11


NO/cGMP/PKG signal<strong>in</strong>g mediate the migration <strong>of</strong> hNPCs<br />

Several studies implicate the <strong>in</strong>volvement <strong>of</strong> NO-cGMP signal<strong>in</strong>g <strong>in</strong> neuronal migration dur<strong>in</strong>g<br />

nervous system development (Tanaka et al., 1994; Haase and Bicker, 2003; Peunova et al., 2007;<br />

Knipp and Bicker, 2009) and after CNS <strong>in</strong>jury <strong>in</strong> adults (Zhang et al., 2001; Cui et al., 2009;<br />

Gutierrez-Mec<strong>in</strong>as et al., 2007). In this study we provide the first experimental evidence that NOcGMP<br />

signal<strong>in</strong>g positively regulate the migration <strong>of</strong> fetal human neuronal progenitor cells. Firstly,<br />

we showed expression <strong>of</strong> NO-sensitive sGC <strong>in</strong> cells that migrate out <strong>of</strong> human neurospheres by<br />

immunocytochemical detection <strong>of</strong> cGMP. Upon stimulation with NO donor the number <strong>of</strong> cGMP-<br />

IR cells <strong>in</strong>creased significantly, which was reduced <strong>in</strong> the presence <strong>of</strong> sGC <strong>in</strong>hibitor. Thus, our data<br />

clearly demonstrate for the first time to our knowledge that NO-sensitive sGC is present early<br />

dur<strong>in</strong>g human bra<strong>in</strong> development. Co-sta<strong>in</strong><strong>in</strong>g <strong>of</strong> cGMP with cytoskeletal markers <strong>in</strong>dicated that the<br />

major proportion <strong>of</strong> the cGMP produc<strong>in</strong>g cells were GFAP-IR with a lower percentage <strong>of</strong> nest<strong>in</strong><br />

and βIII-tubul<strong>in</strong>-IR cells. A close exam<strong>in</strong>ation <strong>of</strong> the forefront <strong>of</strong> migration revealed ma<strong>in</strong>ly nest<strong>in</strong><br />

and GFAP positive cells which are double-labeled for cGMP (Fig. 3 E-G). These migratory nest<strong>in</strong><br />

and GFAP positive cells are trailed by βIII-tubul<strong>in</strong>-IR neuronal cells (Fig. 3 C & F).<br />

Application <strong>of</strong> enzyme <strong>in</strong>hibitors <strong>of</strong> the sGC/PKG pathway <strong>in</strong> hNPCs culture significantly reduced<br />

the migration <strong>of</strong> cells out <strong>of</strong> the neurospheres (Fig. 4 A-C). This loss <strong>of</strong> function effect and the<br />

presence <strong>of</strong> cGMP-IR cells at the forefront <strong>of</strong> migration (Fig. 3 E-G) <strong>in</strong>fer that a certa<strong>in</strong> level <strong>of</strong><br />

cGMP is required to facilitate the migration <strong>of</strong> cells out <strong>of</strong> the neurospheres. Morphological<br />

observation <strong>of</strong> neurospheres and LDH assay revealed no sign <strong>of</strong> cytotoxicity caused by sGC/PKG<br />

<strong>in</strong>hibitors. Moreover, a cell membrane permeable analogue <strong>of</strong> cGMP, 8-Br-cGMP partially rescued<br />

the block<strong>in</strong>g <strong>of</strong> cell migration by ODQ (Fig. 4 A, D). Thus, it is unlikely that unspecific side effects<br />

<strong>of</strong> the chemical blocker contribute to <strong>in</strong>hibition <strong>of</strong> cell migration.<br />

In a ga<strong>in</strong> <strong>of</strong> function experiments, exposure <strong>of</strong> hNPCs to the NO donor NOC-18 (1-10 µM) and<br />

cGMP analogue (8-Br-cGMP) significantly facilitated the migration <strong>of</strong> cells out <strong>of</strong> the neurospheres<br />

(Fig. 5 A, B, E). This ga<strong>in</strong> <strong>of</strong> function experiment is <strong>in</strong> l<strong>in</strong>e with our recent f<strong>in</strong>d<strong>in</strong>g on the ret<strong>in</strong>oic<br />

acid <strong>in</strong>duced differentiat<strong>in</strong>g human neuronal precursor (NT2) cells (Tegenge and Bicker, 2009). In<br />

NT2 cells, NOC-18 facilitated cell motility <strong>in</strong> a concentration range <strong>of</strong> 1-100 µM, above which we<br />

observed deleterious effect on cell viability and proliferation (Tegenge and Bicker, 2009). Here, a<br />

concentration <strong>of</strong> 100 µM <strong>of</strong> NOC-18 significantly <strong>in</strong>hibited cell migration (Fig. 5 A, B) without<br />

alter<strong>in</strong>g cell viability (Fig. 5 C). The <strong>in</strong>verted U-shaped dose-response curve (Fig. 4B) suggests that<br />

NO is play<strong>in</strong>g a dual role, result<strong>in</strong>g <strong>in</strong> a facilitation <strong>of</strong> cell migration at low concentration while<br />

slow<strong>in</strong>g down at higher concentrations. Similar concentration dependent effects on growth cone<br />

12


motility have been observed on develop<strong>in</strong>g snail neurites (Trimm and Rehder, 2004) and migrat<strong>in</strong>g<br />

neutrophils (Elfer<strong>in</strong>k and VanUffelen, 1996).<br />

To directly l<strong>in</strong>k the effect <strong>of</strong> NO with its potential downstream target enzymes, NO donor was co-<br />

applied with sGC or PKG <strong>in</strong>hibitors. A significant reduction <strong>of</strong> NO-<strong>in</strong>duced facilitation <strong>of</strong> cell<br />

migration was observed when NOC-18 was used <strong>in</strong> comb<strong>in</strong>ation with ODQ or Rp-8-Br-cGMP (Fig.<br />

5 D). This provides further evidence for the <strong>in</strong>volvement <strong>of</strong> cGMP and PKG <strong>in</strong> NO mediated<br />

neuronal progenitor cell migration. The detailed mechanism by which cGMP activated PKG<br />

facilitates cell motility is presently unknown. In both neuronal and non-neuronal cells<br />

reorganization <strong>of</strong> act<strong>in</strong> cytoskeleton is caused downstream <strong>of</strong> NO/cGMP pathway (Brown et al.,<br />

1999; Haase and Bicker 2003, Borán and García, 2007), which possibly <strong>in</strong>volve <strong>in</strong>hibition <strong>of</strong> RhoA<br />

GTPase (Sawada et al. 2001; Gudi et al. 2002; Borán and García 2007). The PKG substrate<br />

Enabled/vasodilator-stimulated phosphoprote<strong>in</strong> (Ena/VASP) family prote<strong>in</strong>s that regulate act<strong>in</strong><br />

polymerization have been also suggested to mediate the action <strong>of</strong> cGMP dur<strong>in</strong>g cell motility<br />

(Sporbert et al., 1999; Chen et al., 2008; L<strong>in</strong>dsay et al., 2007).<br />

A concentration gradient <strong>of</strong> endogenously produced NO has been proposed to regulate cell-cycle<br />

progression <strong>in</strong> the develop<strong>in</strong>g chicken neural tube (Traister et al., 2002). Moreover, dur<strong>in</strong>g early<br />

development <strong>of</strong> Xenopus NO has been shown to differentially coord<strong>in</strong>ate cell proliferation and<br />

motility (Peunova et al., 2007). Likewise, <strong>in</strong> the <strong>in</strong>tact fetal human bra<strong>in</strong> gradients <strong>of</strong> the diffusible<br />

messenger NO may be essential to coord<strong>in</strong>ate the dist<strong>in</strong>ct processes <strong>of</strong> cell proliferation,<br />

differentiation and motility.<br />

Acknowledgments<br />

We thank Dr J. de Vente for his k<strong>in</strong>d gift <strong>of</strong> the cGMP antiserum, Dr. M. Stern for the discussion<br />

dur<strong>in</strong>g the preparation <strong>of</strong> the manuscript and Nad<strong>in</strong>e Seiferth for technical support. M.A. Tegenge<br />

received a Georg-Christoph-Lichtenberg scholarship from the M<strong>in</strong>istry for Science and Culture <strong>of</strong><br />

Lower Saxony. This work was supported by the BMBF grants (013925C) to E. Fritsche, (013925D)<br />

to G. Bicker and DFG grant (FG 1103, BI 262/16-1).<br />

References<br />

Ayala R, Shu T, Tsai L (2007) Trekk<strong>in</strong>g across the bra<strong>in</strong>: the journey <strong>of</strong> neuronal migration. Cell<br />

128: 29-43.<br />

Bicker G (2005) STOP and GO with NO: nitric oxide as a regulator <strong>of</strong> cell motility <strong>in</strong> simple<br />

bra<strong>in</strong>s. Bioessays 27: 495-505.<br />

Brannen CL, Sugaya K (2000) In vitro differentiation <strong>of</strong> multipotent human neural progenitors <strong>in</strong><br />

serum-free medium. Neuroreport 11: 1123-8.<br />

13


Bredt, D. S. and Snyder, S. H. (1994) Transient nitric oxide synthase neurons <strong>in</strong> embryonic cerebral<br />

cortical plate, sensory ganglia, and olfactory epithelium. Neuron 13, 301-313.<br />

Chen H., Lev<strong>in</strong>e Y.C., Golan D.E., Michel T. and L<strong>in</strong> A.J. (2008) ANP-<strong>in</strong>itiated cGMP pathways<br />

regulate VASP phosphorylation and angiogenesis <strong>in</strong> vascular endothelium. J. Biol. Chem. 283,<br />

4439-4447.<br />

Cui X, Chen J, Zacharek A, Roberts C, Yang Y, Chopp M (2009) <strong>Nitric</strong> oxide donor up-regulation<br />

<strong>of</strong> SDF1/CXCR4 and Ang1/Tie2 promotes neuroblast cell migration after stroke. J Neurosci Res 87:<br />

86-95.<br />

Cárdenas A, Moro MA, Hurtado O, Leza JC, Lizasoa<strong>in</strong> I (2005) Dual role <strong>of</strong> nitric oxide <strong>in</strong> adult<br />

neurogenesis. Bra<strong>in</strong> Res Bra<strong>in</strong> Res Rev 50: 1-6.<br />

De Vente J, Ste<strong>in</strong>busch HW, Schipper J (1987) A new approach to immunocytochemistry <strong>of</strong> 3', 5’cyclic<br />

guanos<strong>in</strong>e monophosphate: preparation, specificity, and <strong>in</strong>itial application <strong>of</strong> a new antiserum<br />

aga<strong>in</strong>st formaldehyde-fixed 3', 5’-cyclic guanos<strong>in</strong>e monophosphate. Neuroscience 22: 361-73.<br />

Elfer<strong>in</strong>k JG, VanUffelen BE (1996) The role <strong>of</strong> cyclic nucleotides <strong>in</strong> neutrophil migration. Gen<br />

Pharmaco 27: 387-393.<br />

Enikolopov G, Banerji J, Kuz<strong>in</strong> B (1999) <strong>Nitric</strong> oxide and Drosophila development. Cell Death Differ<br />

6: 956-63.<br />

Foster JA, Phelps PE (2000) Neurons express<strong>in</strong>g NADPH-diaphorase <strong>in</strong> the develop<strong>in</strong>g human<br />

sp<strong>in</strong>al cord. J Comp Neurol 427: 417-27.<br />

Fritsche E, Cl<strong>in</strong>e JE, Nguyen NH, Scanlan TS, Abel J (2005) Polychlor<strong>in</strong>ated biphenyls disturb<br />

differentiation <strong>of</strong> normal human neural progenitor cells: clue for <strong>in</strong>volvement <strong>of</strong> thyroid hormone<br />

receptors. Environ Health Perspect 113: 871-6.<br />

Garthwaite J (2008) Concepts <strong>of</strong> neural nitric oxide-mediated transmission. Eur J Neurosci 27: 2783-<br />

802.<br />

Gudi T., Chen J.C., Casteel D.E., Seasholtz T.M., Boss G.R. and Pilz R.B. (2002) cGMP-dependent<br />

prote<strong>in</strong> k<strong>in</strong>ase <strong>in</strong>hibits serum-response element-dependent transcription by <strong>in</strong>hibit<strong>in</strong>g rho activation<br />

and functions. J. Biol. Chem. 277, 37382-93.<br />

Gutièrrez-Mec<strong>in</strong>as M, Crespo C, Blasco-Ibáñez JM, Nácher J, Varea E, Martínez-Guijarro FJ<br />

(2007) Migrat<strong>in</strong>g neuroblasts <strong>of</strong> the rostral migratory stream are putative targets for the action <strong>of</strong><br />

nitric oxide. Eur J Neurosci 26: 392-402.<br />

Haase A, Bicker G (2003) <strong>Nitric</strong> oxide and cyclic nucleotides are regulators <strong>of</strong> neuronal migration<br />

<strong>in</strong> an <strong>in</strong>sect embryo. Development 130: 3977-3987.<br />

Knipp S, Bicker G (2009) Regulation <strong>of</strong> enteric neuron migration by the gaseous messenger<br />

molecules CO and NO. Development 136: 85-93.<br />

14


L<strong>in</strong>dsay, S.L., Ramsey, S., Aitchison, M., Renné, T. and Evans, T.J. (2007). Modulation <strong>of</strong><br />

lamellipodial structure and dynamics by NO-dependent phosphorylation <strong>of</strong> VASP Ser239. J. Cell<br />

Sci. 120, 3011-3021.<br />

Moors M, Cl<strong>in</strong>e JE, Abel J, Fritsche E (2007) ERK-dependent and <strong>in</strong>dependent pathways trigger<br />

human neural progenitor cell migration. Toxicol Appl Pharmaco 221: 57-67.<br />

Moors M, Rockel TD, Abel J, Cl<strong>in</strong>e JE, Gassmann, K, Schreiber T, Schuwald J, We<strong>in</strong>mann, N,<br />

Fritsche E (2009) <strong>Human</strong> neurospheres as three-dimensional cellular systems for developmental<br />

neurotoxicity test<strong>in</strong>g. Environ Health Perspect 117: 1131-8.<br />

Moreno-López B, Noval JA, González-Bonet LG, Estrada C. (2000) Morphological bases for a role<br />

<strong>of</strong> nitric oxide <strong>in</strong> adult neurogenesis. Bra<strong>in</strong> Res 869: 244-250.<br />

Nott, A., Watson, P. M., Rob<strong>in</strong>son, J. D., Crepaldi, L., and Riccio, A. (2008). S-Nitrosylation <strong>of</strong><br />

histone deacetylase 2 <strong>in</strong>duces chromat<strong>in</strong> remodell<strong>in</strong>g <strong>in</strong> neurons. Nature 455, 411-415.<br />

Peunova N, Sche<strong>in</strong>ker V, Ravi K, Enikolopov G (2007) <strong>Nitric</strong> oxide coord<strong>in</strong>ates cell proliferation<br />

and cell movements dur<strong>in</strong>g early development <strong>of</strong> Xenopus. Cell Cycle 6: 3132-44.<br />

Sawada N., Itoh H., Yamashita J., Doi K., Inoue M., Masatsugu K., Fukunaga Y., Sakaguchi S.,<br />

Sone M., Yamahara K., Yurugi T. and Nakao K. (2001) cGMP-dependent prote<strong>in</strong> k<strong>in</strong>ase<br />

phosphorylates and <strong>in</strong>activates RhoA. Biochem. Biophys. Res. Commun. 280, 798-805.<br />

Sporbert, A., Mertsch, K., Smolenski, A., Hasel<strong>of</strong>f, R.F., Schönfelder, G., Paul, M., Ruth, P.,<br />

Walter, U. and Blasig, I.E. (1999). Phosphorylation <strong>of</strong> vasodilator-stimulated phosphoprote<strong>in</strong>: a<br />

consequence <strong>of</strong> nitric oxide- and cGMP-mediated signal transduction <strong>in</strong> bra<strong>in</strong> capillary endothelial<br />

cells and astrocytes. Bra<strong>in</strong> Res. Mol. Bra<strong>in</strong> Res. 67, 258-266.<br />

Tanaka M, Yoshida S, Yano M, Hanaoka F (1994) Roles <strong>of</strong> endogenous nitric oxide <strong>in</strong> cerebellar<br />

cortical development <strong>in</strong> slice cultures. Neuroreport 5: 2049-2052.<br />

Tegenge MA, Bicker G (2009) <strong>Nitric</strong> oxide and cGMP signal transduction positively regulates the<br />

motility <strong>of</strong> human neuronal precursor (NT2) cells. J Neurochem 110:1828-1841.<br />

Traister A, Abashidze S, Gold V, Plachta N, Karchovsky E, Patel K, Weil M (2002) Evidence that<br />

nitric oxide regulates cell-cycle progression <strong>in</strong> the develop<strong>in</strong>g chick neuroepithelium. Dev Dyn 225:<br />

271-6.<br />

Trimm KR, Rehder V (2004) <strong>Nitric</strong> oxide acts as a slow-down and search signal <strong>in</strong> develop<strong>in</strong>g<br />

neurites. Eur J Neurosci 19: 809-18.<br />

Zhang R, Zhang L, Zhang Z, Wang Y, Lu M, Lapo<strong>in</strong>te M, Chopp M (2001) A nitric oxide donor<br />

<strong>in</strong>duces neurogenesis and reduces functional deficits after stroke <strong>in</strong> rats. Ann Neurol 50: 602-611.<br />

15


Declaration<br />

I herewith declare that I autonomously carried out the PhD-thesis entitled<br />

“<strong>Nitric</strong> <strong>Oxide</strong> <strong>Mediated</strong> <strong>Signal</strong> <strong>Transduction</strong> <strong>in</strong> <strong>Networks</strong> <strong>of</strong> <strong>Human</strong> Model Neurons”.<br />

I did not receive any assistance <strong>in</strong> return for payment by consult<strong>in</strong>g agencies or any other<br />

person. No one received any k<strong>in</strong>d <strong>of</strong> payment for direct or <strong>in</strong>direct assistance <strong>in</strong> correlation to<br />

the content <strong>of</strong> the submitted thesis.<br />

I conducted the project at the follow<strong>in</strong>g <strong>in</strong>stitutions:<br />

Division <strong>of</strong> Cell Biology, Institute <strong>of</strong> Physiology, University <strong>of</strong> Veter<strong>in</strong>ary Medic<strong>in</strong>e<br />

Hannover, Germany.<br />

The thesis has not been submitted elsewhere for an exam, as thesis or for evaluation <strong>in</strong> a<br />

similar context.<br />

I hereby affirm the above statements to be complete and true to the best <strong>of</strong> my knowledge.<br />

_______________________________<br />

Million Adane Tegenge

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!