DOI QR코드

DOI QR Code

Elevated RalA activity in the hippocampus of PI3Kγ knock-out mice lacking NMDAR-dependent long-term depression

  • Sim, Su-Eon (Department of Brain and Cognitive Sciences, College of Natural Sciences, Seoul National University) ;
  • Lee, Hye-Ryeon (National Creative Research Initiative Center for Memory, Department of Biological Sciences, College of Natural Sciences, Seoul National University) ;
  • Kim, Jae-Ick (National Creative Research Initiative Center for Memory, Department of Biological Sciences, College of Natural Sciences, Seoul National University) ;
  • Choi, Sun-Lim (National Creative Research Initiative Center for Memory, Department of Biological Sciences, College of Natural Sciences, Seoul National University) ;
  • Bakes, Joseph (Department of Brain and Cognitive Sciences, College of Natural Sciences, Seoul National University) ;
  • Jang, Deok-Jin (Department of Applied Biology, College of Ecology and Environment, Kyungpook National University) ;
  • Lee, Kyungmin (Department of Anatomy, School of Medicine, Brain Science & Engineering Institute, Kyungpook National University) ;
  • Han, Kihoon (Department of Biological Sciences, KAIST) ;
  • Kim, Eunjoon (Department of Biological Sciences, KAIST) ;
  • Kaang, Bong-Kiun (Department of Brain and Cognitive Sciences, College of Natural Sciences, Seoul National University)
  • Received : 2012.07.12
  • Accepted : 2012.09.12
  • Published : 2013.02.28

Abstract

Phosphoinositide 3-kinases (PI3Ks) play key roles in synaptic plasticity and cognitive functions in the brain. We recently found that genetic deletion of $PI3K{\gamma}$, the only known member of class IB PI3Ks, results in impaired N-methyl-D-aspartate receptor-dependent long-term depression (NMDAR-LTD) in the hippocampus. The activity of RalA, a small GTP-binding protein, increases following NMDAR-LTD inducing stimuli, and this increase in RalA activity is essential for inducing NMDAR-LTD. We found that RalA activity increased significantly in $PI3K{\gamma}$ knockout mice. Furthermore, NMDAR-LTD-inducing stimuli did not increase RalA activity in $PI3K{\gamma}$ knockout mice. These results suggest that constitutively increased RalA activity occludes further increases in RalA activity during induction of LTD, causing impaired NMDAR-LTD. We propose that $PI3K{\gamma}$ regulates the activity of RalA, which is one of the molecular mechanisms inducing NMDAR-dependent LTD.

Keywords

INTRODUCTION

Phosphoinositide 3-kinases (PI3Ks) are a family of enzymes involved in numerous cellular functions such as cell growth, survival, proliferation, migration, and intracellular vesicular transport (1-3). PI3Ks have unique and complex kinase properties, with dual-kinase activity comprised of both lipid kinase activity, transferring phosphate to the 3-position of the inositol ring in phosphoinositide substrates, and protein kinase activity, transferring phosphate to specific Ser/Thr on protein substrates (4,5). PI3Ks have been classified into three major classes such as class I, class II, and class III based on their structural features and functional similarities. Class I PI3Ks are further subdivided into classes IA and IB according to the composition of the catalytic and regulatory subunits constituting the PI3K heterodimers (1,2,6). The roles of PI3Ks in synaptic plasticity of receptor trafficking, fear memory, memory retrieval and memoryf extinction in the brain have been studied extensively (7-14). In contrast to previous studies that have focused on broad ranges of PI3K activities and used general PI3K inhibitors such as wortmannin and LY294002, our group recently reported the specific roles of PI3Kγ, the only known member of class IB PI3Ks (15). In that study, we reported that genetic deletion or pharmacological inhibition of PI3Kγ impairs N-methyl-D-aspartate receptor-dependent long-term depression (NMDARLTD) and behavioral flexibility. We also revealed that p38 mitogen- activated protein kinase (MAPK), but not glycogen synthase kinase 3 beta or Akt, is involved in PI3Kγ-mediated NMDAR-LTD.

Ras-like protein A (RalA) is a well-known small GTPase involved in membrane trafficking such as endocytosis and exocytosis and is thought to have a role in NMDAR-LTD (16-18). In particular, a study showed that NMDAR activation by NMDA in cultured neurons increases RalA activity (17). Furthermore, PI3Ks regulate RalA activation via the Rap1/PI3K/RalGEFs pathway (19-23).

As NMDAR-LTD is a complex cellular event involving dozens of molecules, we proposed that other molecules, besides p38 MAPK, might have a role in PI3Kγ-mediated NMDARLTD. In this study, we examined the activity of RalA as a putative downstream molecule in PI3Kγ-mediated NMDAR-LTD.

 

RESULTS AND DISCUSSION

We performed a RalA pull-down assay after applying an NMDAR-LTD-inducing stimulus to hippocampal slices from PI3Kγ knock-out (KO) mice and their wild-type littermates to investigate changes in RalA activity (24). We delivered a low-frequency stimulus (LFS; 1 Hz, 900 s) to acute hippocampal slices, submitting only the CA1 region of each slice to the RalA pull-down assay. We reported previously that NMDAR-LTD is robustly induced by LFS under our experiment conditions (15). The GTP-bound active form of RalA specifically binds to GST-RalBD. In accordance with previous studies, the proportion of RalA in its active form increased significantly by LFS treatment in wild-type littermates (17). However, without LFS, RalA activity increased in PI3Kγ KO mice in comparison to that in wild-type littermates, and LFS induced no further increase in RalA activity (Fig. 1). This result suggests that genetic deletion of PI3Kγ causes increased RalA activity and impedes the activity-dependent increase in RalA activity by LFS. We conclude from these results that the impaired NMDAR-LTD in PI3Kγ KO mice is caused by excessive RalA activity.

We found abnormally elevated RalA activity in PI3Kγ KO mice using electrophysiology and biochemistry. Unlike wild-type littermates, the LFS stimulus failed to induce increases in RalA activity in PI3Kγ KO mice. We propose that elevated RalA activity in PI3Kγ KO mice may impair induction of NMDAR-LTD.

The main question arising from our results concerns the mechanism underlying the increase in basal RalA activity in PI3Kγ KO mice. Previous studies have shown that RalA is activated by cAMP-mediated activation of protein kinase A in epithelial cells (25,26). This implies that increases in cAMP might increase RalA activity. PI3Kγ has been extensively studied in the heart as a key molecule regulating cardiac contractility. PI3Kγ negatively regulates the cAMP level in cardiomyocytes via direct interaction with phosphodiesterase 3B (27-29). Based on these results, we might assume that basal cAMP levels would increase in PI3Kγ KO mice brains, as well as their hearts, leading to increased RalA activity.

Fig. 1.RalA activation by N-methyl-d-aspartate receptor-dependent long-term depression (NMDAR-LTD) stimulus in phosphoinositide 3-kinase (PI3K)γ knockout (KO) and wild-type littermates. (A) A representative Western blotting image. RalA activity changes in CA1 regions after the NMDAR-LTD stimulus in hippocampi of wild-type littermates and PI3Kγ KO mice. (B) Quantification of active/total RalA ratio. The ratio of active/total RalA was normalized to that of the wild-type littermate control group (n = 3). The stimulus-by-genotyping interaction was statistically significant as assessed by two-way analysis of variance (F1,8 = 12.92, P < 0.01). Post-hoc analysis using Tukey’s test revealed significant differences between the control and low-frequency stimulus (LFS) group within the wild-type (**P < 0.01) and between wild-type littermates and PI3Kγ KO mice within the control group (***P < 0.001).

Another important question is how the basally occluded RalA activity prevents the induction of NMDAR-LTD in PI3Kγ KO mice. A previous study showed that activated RalA, mainly induced by activating NMDAR, directly binds to Ral-binding protein 1 (RalBP1) and promotes translocation of both proteins to dendritic spines, which is essential for NMDAR-LTD (17). We assumed that the increase in RalA activity via NMDAR activation is important to trigger the translocation of RalBP1 to dendritic spines. The absence of an activity-dependent increase in RalA activity in PI3Kγ KO mice may fail to recruit RalBP1 to synapses, which causes impairment of NMDARLTD. Furthermore, previously activated RalA without NMDAR activation may disrupt recruitment of other components critical for LTD induction. Further studies are required to elucidate the precise mechanisms of altered RalA activity inPI3KγKO mice.

The occluded RalA activity in PI3Kγ KO mice suggests that occlusion of NMDAR-LTD occurs via reduced AMPAR levels in the basal-state. However, our previous study showed that basal synaptic transmission of PI3Kγ KO mice is normal, indicating that occluded RalA activity has no effect on AMPAR levels in the basal state (15). This is also supported by another study showing that overexpression of the constitutive active form of RalA does not change basal AMPAR levels in cultured neurons (17).

Taken together, our study showed that genetic deletion of PI3Kγ causes an elevation in RalA activity, which might be one of the main reasons for NMDAR-LTD impairment in PI3Kγ KO mice. These findings also support a novel functional relationship between PI3Kγ and RalA, two key molecules involved in NMDAR-LTD.

 

MATERIALS AND METHODS

Animals

The care and information of Pik3cg-/- and wild-type littermates have been described previously (15,30). Animals were cared for in accordance with the regulations and guidelines of the Animal Care and Use Committee of Seoul National University

Sample preparation

Transverse hippocampal slices (400-μm thick) were prepared from 4-5-week-old PI3Kγ wild-type littermates and KO mice using a manual tissue chopper. Slices were recovered in an interface chamber for at least 2 h before applying LFS-LTD (900 stimuli, 1 Hz). The interface chamber was maintained at 28℃ and consistently supplied with oxygenated artificial cerebrospinal fluid containing 124 mM NaCl, 2.5 mM KCl, 1 mM NaH2PO4, 25 mM NaHCO3, 10 mM glucose, 2 mM CaCl2, and 2 mM MgSO4. The intensity of the LTD stimulus was adjusted to approximately 40% of the maximum field excitatory postsynaptic potential slopes. Each slice was incubated for an additional 20 min following delivery of LFS-LTD, and the CA1 regions from each slice were immediately frozen in liquid nitrogen.

RalA pull-down assay with stimulated slices

The CA1 regions of stimulated hippocampal slices from individual mice were pooled to obtain 300-400 μg of protein from each mouse. The RalA pull-down assay was performed as described previously (24). Briefly, 40 μg of GST-RalBD was preincubated with 35 μl of glutathione-Sepharose beads at 4℃ for 1 h and washed three times. Sample tissue was lysed in Ral buffer (10% glycerol, 1% Nonidet P-40, 50 mM Tris-HCl, 200 mM NaCl, and 10 mM MgCl2) with protease inhibitor cocktail (Roche, Manheim, Germany) and phosphatase inhibitor cocktails I and II (Sigma, St. Louis, MO, USA), and lysates were centrifuged at 16,000 × g for 20 min at 4℃. The supernatant was added to GST-RalBD pre-coupled beads and incubated for 4 h at 4℃. The incubated beads were washed four times in Ral buffer, and surplus buffer was removed with a 1-ml syringe. Proteins were eluted with sodium dodecyl sulfate (SDS) sample buffer and resolved on SDS-polyacrylamide gel electrophoresis. The amount of proteins was measured by Western blotting using the following antibodies: anti-RalA (1:5,000; BD Biosciences, San Diego, CA, USA), anti-actin (1:10,000; Sigma), and anti-GST (1:10,000; Bethyl Laboratories, Montgomery, TX, USA).

References

  1. Hawkins, P., Anderson, K., Davidson, K. and Stephens, L. (2006) Signalling through Class I PI3Ks in mammalian cells. Biochem. Soc. Trans. 34, 647-662. https://doi.org/10.1042/BST0340647
  2. Vanhaesebroeck, B., Guillermet-Guibert, J., Graupera, M. and Bilanges, B. (2010) The emerging mechanisms of isoform- specific PI3K signalling. Nat. Rev. Mol. Cell Bio. 11, 329-341. https://doi.org/10.1038/nrm2882
  3. Han, S. J. and Choi, K. Y. (2010) Novel p104 protein regulates cell proliferation through PI3K inhibition and p27Kip1 expression. BMB Rep. 43, 199-204. https://doi.org/10.5483/BMBRep.2010.43.3.199
  4. Bondeva, T., Pirola, L., Bulgarelli-Leva, G., Rubio, I., Wetzker, R. and Wymann, M. P. (1998) Bifurcation of lipid and protein kinase signals of PI3K${\gamma}$ to the protein kinases PKB and MAPK. Science 282, 293. https://doi.org/10.1126/science.282.5387.293
  5. Ruckle, T., Schwarz, M. K. and Rommel, C. (2006) PI3K${\gamma}$ inhibition: towards an'aspirin of the 21st century'? Nat. Rev. Drug Discov. 5, 903-918. https://doi.org/10.1038/nrd2145
  6. Kok, K., Geering, B. and Vanhaesebroeck, B. (2009) Regulation of phosphoinositide 3-kinase expression in health and disease. Trends Biochem. Sci. 34, 115-127. https://doi.org/10.1016/j.tibs.2009.01.003
  7. Lin, C. H., Yeh, S. H., Lu, K. T., Leu, T. H., Chang, W. C. and Gean, P. W. (2001) A role for the PI-3 kinase signaling pathway in fear conditioning and synaptic plasticity in the amygdala. Neuron 31, 841-851. https://doi.org/10.1016/S0896-6273(01)00433-0
  8. Sanna, P. P., Cammalleri, M., Berton, F., Simpson, C., Lutjens, R., Bloom, F. E. and Francesconi, W. (2002) Phosphatidylinositol 3-kinase is required for the expression but not for the induction or the maintenance of long-term potentiation in the hippocampal CA1 region. J. Neurosci. 22, 3359-3365.
  9. Man, H. Y., Wang, Q., Lu, W. Y., Ju, W., Ahmadian, G., Liu, L., D'Souza, S., Wong, T., Taghibiglou, C., Lu, J., Becker, L. E., Pei, L., Liu, F., Wymann, M. P., MacDonald, J. P. and Wang, Y. T. (2003) Activation of PI3-kinase is required for AMPA receptor insertion during LTP of mEPSCs in cultured hippocampal neurons. Neuron 38, 611-624. https://doi.org/10.1016/S0896-6273(03)00228-9
  10. Opazo, P., Watabe, A. M., Grant, S. G. N. and O'Dell, T. J. (2003) Phosphatidylinositol 3-kinase regulates the induction of long-term potentiation through extracellular signal-related kinase-independent mechanisms. J. Neurosci. 23, 3679-3688.
  11. Chen, X., Garelick, M. G., Wang, H., Li, V., Athos, J. and Storm, D. R. (2005) PI3 kinase signaling is required for retrieval and extinction of contextual memory. Nat. Neuro sci. 8, 925-931. https://doi.org/10.1038/nn1482
  12. Peineau, S., Taghibiglou, C., Bradley, C., Wong, T. P., Liu, L., Lu, J., Lo, E., Wu, D., Saule, E., Bouschet, T., Matthews, P., Isaac, J. T. R., Bortolotto, Z. A., Wang, Y. T. and Collingridge, G. L. (2007) LTP inhibits LTD in the hippocampus via regulation of GSK3${\beta}$. Neuron 53, 703-717. https://doi.org/10.1016/j.neuron.2007.01.029
  13. Arendt, K. L., Royo, M., Fernandez-Monreal, M., Knafo, S., Petrok, C. N., Martens, J. R. and Esteban, J. A. (2009) PIP3 controls synaptic function by maintaining AMPA receptor clustering at the postsynaptic membrane. Nat. Neurosci. 13, 36-44.
  14. Peineau, S., Nicolas, C. S., Bortolotto, Z. A., Bhat, R. V., Ryves, W. J., Harwood, A. J., Dournaud, P., Fitzjohn, S. M. and Collingridge, G. L. (2009) A systematic investigation of the protein kinases involved in NMDA receptor-dependent LTD: evidence for a role of GSK-3 but not other serine/threonine kinases. Mol. Brain 2, 22. https://doi.org/10.1186/1756-6606-2-22
  15. Kim, J. I., Lee, H. R., Sim, S., Baek, J., Yu, N. K., Choi, J. H., Ko, H. G., Lee, Y. S., Park, S. W. Kwak, C., Ahn, S. J., Choi, S. Y., Kim, H., Kim, K. H., Backx, P. H., Bradley, C. A., Kim, E., Jang, D. J., Lee, K., Kim, S. J., Zhuo, M., Collingridge, G. L. and Kaang, B. K. (2011) PI3K [gamma] is required for NMDA receptor-dependent long-term depression and behavioral flexibility. Nat. Neurosci. 14, 1447-1454. https://doi.org/10.1038/nn.2937
  16. van Dam, E. M. and Robinson, P. J. (2006) Ral: mediator of membrane trafficking. Int. J. Biochem. Cell Biol. 38, 1841-1847. https://doi.org/10.1016/j.biocel.2006.04.006
  17. Han, K., Kim, M. H., Seeburg, D., Seo, J., Verpelli, C., Han, S., Chung, H. S., Ko, J., Lee, H. W., Kim, K., Heo, W., Meyer, T., Kim, H., Sala, C., Choi, S. Sheng, M. and Kim, E. (2009) Regulated RalBP1 binding to RalA and PSD-95 controls AMPA receptor endocytosis and LTD. PLoS Biol. 7, e1000187. https://doi.org/10.1371/journal.pbio.1000187
  18. Collingridge, G. L., Peineau, S., Howland, J. G. and Wang, Y. T. (2010) Long-term depression in the CNS. Nat. Rev. Neurosci. 11, 459-473.
  19. Bos, J. L., de Rooij, J. and Reedquist, K. A. (2001) Rap1 signalling: adhering to new models. Nat. Rev. Mol. Cell Bio. 2, 369-377. https://doi.org/10.1038/35073073
  20. Chen, X. W., Leto, D., Chiang, S. H., Wang, Q. and Saltiel, A. R. (2007) Activation of RalA is required for insulin-stimulated Glut4 trafficking to the plasma membrane via the exocyst and the motor protein Myo1c. Dev. Cell 13, 391-404. https://doi.org/10.1016/j.devcel.2007.07.007
  21. Tian, X., Rusanescu, G., Hou, W., Schaffhausen, B. and Feig, L. A. (2002) PDK1 mediates growth factor-induced Ral-GEF activation by a kinase-independent mechanism. Science's STKE 21, 1327.
  22. Feig, L. A. (2003) Ral-GTPases: approaching their 15 minutes of fame. Trends Cell Biol. 13, 419-425. https://doi.org/10.1016/S0962-8924(03)00152-1
  23. Ye, X. and Carew, T. J. (2010) Small G protein signaling in neuronal plasticity and memory formation: the specific role of ras family proteins. Neuron 68, 340-361. https://doi.org/10.1016/j.neuron.2010.09.013
  24. Wolthuis, R. M. F., Franke, B., Van Triest, M., Bauer, B., Cool, R. H., Camonis, J. H., Akkerman, J. W. N. and Bos, J. L. (1998) Activation of the small GTPase Ral in platelets. Mol. Cell. Biol. 18, 2486. https://doi.org/10.1128/MCB.18.5.2486
  25. Rondaij, M. G., Sellink, E., Gijzen, K. A., ten Klooster, J. P., Hordijk, P. L., van Mourik, J. A. and Voorberg, J. (2004) Small GTP-binding protein Ral is involved in cAMP-mediated release of von Willebrand factor from endothelial cells. Arterioscl. Throm. Vas. 24, 1315-1320. https://doi.org/10.1161/01.ATV.0000131267.13425.45
  26. Rondaij, M. G., Bierings, R., Van Agtmaal, E. L., Gijzen, K. A., Sellink, E., Kragt, A., Ferguson, S. S. G., Mertens, K., Hannah, M. J., Van Mourik, J. A., Fernandez-Borja, M. and Voorberg, J. (2008) Guanine exchange factor RalGDS mediates exocytosis of Weibel-Palade bodies from endothelial cells. Blood 112, 56-63. https://doi.org/10.1182/blood-2007-07-099309
  27. Crackower, M. A., Oudit, G. Y., Kozieradzki, I., Sarao, R., Sun, H., Sasaki, T., Hirsch, E., Suzuki, A., Shioi, T., Irie-Sasaki, J., Sah, R., Cheng, H. M., Rybin, V. O., Lembo, G., Fratta, L., Oliveira-dos-Santos, A., Benovic, J. L., Kahn, C. R., Izumo, S., Steinberg, S. F., Wymann, M. P., Backx, P. H. and Penninger, J. M. (2002) Regulation of myocardial contractility and cell size by distinct PI3K-PTEN signaling pathways. Cell 110, 737-749. https://doi.org/10.1016/S0092-8674(02)00969-8
  28. Kerfant, B. G., Rose, R. A., Sun, H. and Backx, P. H. (2006) Phosphoinositide 3-kinase ${\gamma}$ regulates cardiac contractility by locally controlling cyclic adenosine monophosphate levels. Trends Cardiovas. Med. 16, 250-256. https://doi.org/10.1016/j.tcm.2006.04.006
  29. Patrucco, E., Notte, A., Barberis, L., Selvetella, G., Maffei, A., Brancaccio, M., Marengo, S., Russo, G., Azzolino, O., Rybalkin, S. D., Silengo, L., Altruda, F., Wetzker, R., Wymann, M. P., Lembo, G. and Hirsch, E. (2004) PI3Kγ modulates the cardiac response to chronic pressure overload by distinct kinase-dependent and-independent effects. Cell 118, 375-387. https://doi.org/10.1016/j.cell.2004.07.017
  30. Sim, S. E., Park, S. W., Choi, S. L., Yu, N. K., Ko, H. G., Jang, D. J., Lee, K. and Kaang, B. K. (2011) Assessment of the effects of virus-mediated limited Oct4 overexpression on the structure of the hippocampus and behavior in mice. BMB Rep. 44, 793-798. https://doi.org/10.5483/BMBRep.2011.44.12.793