Elsevier

Cellular Signalling

Volume 25, Issue 12, December 2013, Pages 2424-2439
Cellular Signalling

Nitric oxide regulates AKT phosphorylation and nuclear translocation in cultured retinal cells

https://doi.org/10.1016/j.cellsig.2013.08.001Get rights and content

Highlights

  • The nitric oxide donor SNAP promotes AKT phosphorylation in retinal cultures.

  • AKT phosphorylation on Ser473 and Thr308 residues is PKG, PI3K and mTOR-dependent.

  • The nitric oxide synthase substrate l-arginine also stimulates AKT phosphorylation.

  • AKT phosphorylation is upstream-regulated by glutamate and NMDA receptors.

  • Nitric oxide elicits AKT nuclear translocation and cell survival in retinal cultures.

Abstract

Previous studies have shown that nitric oxide (NO) inhibits apoptosis of retinal neurons in culture through the canonical cyclic GMP/protein kinase G (PKG)-dependent pathway, but also involving multiple kinase pathways, such as phosphatidylinositol 3′ kinase (PI3k) and AKT. NO and AKT exhibit survival-promoting properties and display important roles in both CNS development and plasticity. The purpose of this study was to evaluate the effects of exogenous NO, derived from the NO donor S-nitroso-N-acetylpenicillamin (SNAP), or endogenous NO, produced from l-arginine, on AKT phosphorylation in cultured chick retinal neurons. Our results demonstrate that SNAP or l-arginine enhances AKT phosphorylation on both serine-473 and threonine-308 residues in a concentration and time-dependent manner. This effect was mediated by the activation of soluble guanylyl cyclase and PKG, since it was blocked by the respective enzyme inhibitors ODQ or LY83583 and KT5823, as well as by transduction with shRNA lentiviruses coding PKGII shRNA, and mimicked by the respective enzyme activators YC-1 and 8-Bromo cyclic GMP, and also by the cyclic GMP phosphodiesterase inhibitor zaprinast. In addition, LY294002 or wortmannin suppressed the SNAP effect, indicating the involvement of phosphoinositide 3′ kinase. Moreover, the mTOR inhibitor KU0063794 blocked SNAP-induced AKT phosphorylation at both residues, suggesting the participation of the mTORC2 complex in the process. Glutamate and NMDA also promoted AKT phosphorylation and a nitric oxide synthase inhibitor abrogated these effects, revealing a mechanism involving the activation of NMDA receptors and NO production. We have also found that SNAP and l-arginine induced AKT translocation into the nucleus of retinal neurons as well as other neuronal cell lines. SNAP also protects retinal cells from death induced by hydrogen peroxide and this effect was blocked by the phosphoinositide 3′ kinase inhibitor LY294002. We therefore conclude that NO produced from endogenous or exogenous sources promotes AKT activation and its shuttling to the nucleus, probably participating in neuronal survival pathways important during CNS development.

Introduction

Nitric oxide (NO) is synthesized from l-arginine (L-Arg) which is converted to l-citrulline and NO in the presence of O2, NADPH and tetrahydrobiopterin in a nitric oxide synthase (NOS)-catalyzed reaction [1], [2]. Many brain regions are able to synthesize NO [3], although it is most actively formed in the cerebellum [4], [5]. NO is a potent bioactive molecule acting as a diffusible messenger in intercellular communication and intracellular signaling. In postsynaptic neurons, NO is formed following activation of glutamate receptors, mainly the NMDA subtype [6], although recent evidence also indicates the involvement of AMPA receptors [7], [8]. After activation of these receptors, calcium is transiently increased in the cytosol and forms a complex with calmodulin, which binds to and activates neuronal nitric oxide synthase (nNOS) [2]. Previous work has shown the presence of NOS in the vertebrate retina [9], [10]. We have also shown a NADPH diaphorase activity in retinal cells, which could be inhibited by L-Arg analogs and stimulated by calcium during early stages of chick retina development. Moreover, this diaphorase activity was localized in distinct cell types within the retina, including photoreceptors and amacrine cells [11]. Glutamate is able to stimulate NOS activity in the chick retina and this effect is mediated by the activation of NMDA receptors [12], [13], [14]. On the other hand, we have demonstrated the presence of a high affinity system for L-Arg uptake in cultured retinal neurons and glial cells. Furthermore, L-Arg could be released from glial cells and taken up by neurons in these cultures [15]. However, immunocytochemistry for nNOS or l-citruline in mixed neuron/glia cultures as well as measuring nitrite formation in purified neuronal cultures indicated the presence of NOS and NO production in neurons, but not in glial cells [7], [15], [16]. Significantly, we have also shown that the activation of NMDA receptors promoted the inhibition of protein synthesis and, as a consequence, increased the intracellular pool of L-Arg available to be used for the synthesis of NO in cultured retinal cells [17]. NO is also closely involved in the control of glial cell proliferation [18], vitamin C uptake [19] and the modulation of Src, ERK and CREB activation in the retina [7], [8]. Hence, these results indicate the importance of NO in neuronal–glial communication and glutamate/Src/ERK/CREB linkage during retinal development [7], [8], [16], [18].

AKT plays a crucial function in intracellular signaling pathways associated with cell survival and proliferation. AKT activation is usually achieved by a tyrosine kinase receptor-mediated phosphoinositide 3′ kinase (PI3K) activation, which is known to be wortmannin-sensitive [20]. During its activation, AKT is recruited to the plasma membrane where it binds to the PI3K-catalyzed products, PI(3,4,5)P3 and PI(3,4)P2, exposing both Threonine-308 (Thr308) and Serine 473 (Ser473) residues, which in turn can be phosphorylated by 3-phosphoinositide-dependent kinase (PDK1) and mammalian target of rapamycin complex-2 (mTORC2) [21], respectively. This double phosphorylation results in a full activation of AKT [22], [23]. Once activated, AKT induces the phosphorylation of a number of nuclear and cytosolic proteins that regulate cell metabolism, growth and survival. In line with this, recent data showed that genetic depletion of PTEN, a phosphatase that counter-balances PI3K activity and thereby diminishes the downstream activation of AKT, promotes a robust disturbance in retinal neurogenesis [24] through a mechanism related to Notch signaling [25].

Previous studies from our laboratory showed the participation of PI3K in promotion of cell survival by NO in retinal neuronal cultures [16]. Accordingly, herein we present evidence that delineate a novel signaling pathway in which NO, as well as its upstream activating system represented by NMDA receptors, activates AKT. Furthermore, we show that exogenous or endogenous NO elicits AKT translocation to the nucleus, an effect dependent on a PI3K-regulated AKT phosphorylation. NO is also able to protect retinal cells from hydrogen peroxide-induced cell death through a PI3K-dependent pathway. Therefore, the present data highlight the importance of NO in the regulation of survival pathways related to AKT activation in retinal neurons and indicate that NO plays a pivotal role during CNS development.

Section snippets

Reagents

LY294002 and NMDA were from Biomol (Plymouth Meeting, PA, USA). Wortmannin, LY83583, 1-oxo-9.12-epoxy-1H-diidolo[1,2,3-fg:3′,2′,1′kl]pyrrolo[3], [4][1], [6] benzodiazocine-10carboxylic acid methyl ester (KT5823), HEPES, NG-nitro-l-arginine methyl ester (L-NAME), 3-(5′hydroxymethyl-2′-furyl)-1-1benzyl indazole (YC-1), 1H[1], [2], [4]-oxadiazole [4,3-a]quinoxalin-1-one (ODQ), dimethyl sulfoxide (DMSO), diamidino-2-phenylindole, 4′,6-diamidino-2-phenylindole (DAPI), 7-Nitroindazole (7-NI),

Starvation of retinal cultures is essential for lowering basal AKT phosphorylation level

It is well known that serine-threonine kinases can be activated by serum added to the culture medium. The strategy of serum withdrawal was used in different cell cultures and does not cause any cellular alteration or damage but to reduce the basal phosphorylation level of different proteins [28]. Interestingly, starvation of our cultures using serum-free medium for 24 h had no effect on basal AKT phosphorylation when compared with cultures maintained with serum (3% FBS). On the other hand, when

Discussion

NO is able to exert its cellular functions through the activation of the canonical sGC/cGMP/PKG pathway or by protein nitrosylation and nitration [44]. However, very little is known about the signaling pathways triggered by such stimulatory events. Our recent studies revealed that the canonical pathway is involved in the activation of different kinases including ERK as well as the activation of the transcription factor CREB in retinal cells [7], [8]. Here we have studied in cultures of chick

Conclusions

  • SNAP, a NO donor, or L-Arg, the NOS substrate, promotes AKT phosphorylation when added to cultured retinal cells in a concentration and time-dependent manner.

  • The phosphorylation occurs in the residues Ser473 and Thr308, and the phosphorylation at both residues is completely prevented by the PI3K inhibitors LY294002 or wortmannin, or the mTor inhibitor KU0063794.

  • The effect of NO on AKT phosphorylation is mediated by the canonical signaling pathway involving activation of sGC, cGMP production and

Acknowledgments

We thank Luzeli R. de Assis and Sarah A. Rodrigues for the technical assistance. We greatly acknowledge Dr. Edward Ziff for his critical review of the manuscript. This work was supported by grants from CNPq, Capes (Procad 0236055), Faperj and Pronex/Faperj. TAM-G, CCP and TGE were recipients of fellowships from Capes. RPC was a fellow from CNPq and Faperj.

References (63)

  • R. Palmer et al.

    Biochem. Biophys. Res. Commun.

    (1988)
  • D. Grzybicki et al.

    J. Chem. Neuroanat.

    (1996)
  • S.J. East et al.

    Neurosci. Lett.

    (1991)
  • R. Socodato et al.

    J. Biol. Chem.

    (2012)
  • H. Ríos et al.

    Dev. Brain Res.

    (2000)
  • E.J. Lee et al.

    Brain Res.

    (2003)
  • R. Ientile et al.

    FEBS Lett.

    (1996)
  • C.R. Magalhães et al.

    Int. J. Dev. Neurosci.

    (2006)
  • C.C. Portugal et al.

    J. Biol. Chem.

    (2012)
  • D. Brodbeck et al.

    J. Biol. Chem.

    (1999)
  • D.P. Brazil et al.

    Trends Biochem. Sci.

    (2001)
  • K. Sakagami et al.

    Mol. Cell. Neurosci.

    (2012)
  • O.H. Lowry et al.

    J. Biol. Chem.

    (1951)
  • K. Oh-hashi et al.

    Biochem. Biophys. Res. Commun.

    (1999)
  • L. McDonald et al.

    Adv. Pharmacol.

    (1995)
  • A.J. Cristóvão et al.

    Int. J. Dev. Neurosci.

    (2002)
  • A.J. Cristóvão et al.

    Mol. Brain Res.

    (2002)
  • J. Garthwaite et al.

    Eur. J. Pharmacol.

    (1989)
  • J.E. Brenman et al.

    Curr. Opin. Neurobiol.

    (1997)
  • R. Meier et al.

    J. Biol. Chem.

    (1997)
  • A. Martínez-Ruiz et al.

    Free Radic. Biol. Med.

    (2011)
  • A. Contestabile

    Prog. Neurobiol.

    (2008)
  • R. Carini et al.

    Free Radic. Biol. Med.

    (2006)
  • C.-D. Fan et al.

    J. Biol. Chem.

    (2013)
  • A.M. Martelli et al.

    Cell. Signal.

    (2006)
  • R. Knowles et al.

    Biochem. J.

    (1994)
  • D.S. Bredt et al.

    Nature

    (1990)
  • A. Contestabile

    Cerebellum

    (2012)
  • R. Socodato et al.

    J. Neurochem.

    (2009)
  • M.H. de-Faria et al.

    Braz. J. Med. Biol. Res.

    (1995)
  • R. Paes-de-Carvalho et al.

    Rev. Bras. Biol.

    (1996)
  • Cited by (0)

    1

    The first three authors (Telmo A. Mejía-García, Camila C. Portugal, and Thaísa G. Encarnação) have equal contributions to the work.

    View full text