Activity-Dependent Regulation of Neuronal ... - Oxford Journals

3 downloads 0 Views 2MB Size Report
Oct 26, 2007 - Brigadski and Richard Kolarow for help on the organotypic slice cultures. AG and ... Sendtner M, Bellingham MC, Betz H, Noakes PG. 2005.
Cerebral Cortex June 2008;18:1335--1349 doi:10.1093/cercor/bhm165 Advance Access publication October 26, 2007

Activity-Dependent Regulation of Neuronal Apoptosis in Neonatal Mouse Cerebral Cortex

Nicolas Heck1, Antje Golbs1, Therese Riedemann1, Jyh-Jang Sun1, Volkmar Lessmann1,2 and Heiko J. Luhmann1

A massive neuronal loss during early postnatal development has been well documented in the murine cerebral cortex, but the factors that drive cells into apoptosis are largely unknown. The role of neuronal activity in developmental apoptosis was studied in organotypic neocortical slice cultures of newborn mice. Multielectrode array and whole-cell patch-clamp recordings revealed spontaneous network activity characterized by synchronized burst discharges, which could be blocked by tetrodotoxin and ionotropic glutamate receptor antagonists. The identical neuropharmacological manipulations also caused a significant increase in the number of apoptotic neurons as early as 6 h after the start of drug treatment. Moreover, inhibition of the NMDA receptor subunit NR2A or NR2B induced a differential short-term versus delayed increase in the apoptosis rate, respectively. Activation of L-type, voltage-dependent calcium channels was neuroprotective and could prevent activity-dependent apoptosis during NMDA receptor blockade. Furthermore, this effect involved phosphorylation of cAMP response element--binding protein and activation of the tropomyosin-related kinase (Trk) receptors. Inhibition of electrical synapses and blockade of ionotropic g-aminobutyric acid receptors induced specific changes in spontaneous electrical activity patterns, which caused an increase in caspase-3--dependent cell death. Our results demonstrate that synchronized spontaneous network bursts activating ionotropic glutamate receptors promote neuronal survival in the neonatal mouse cerebral cortex.

essential trophic factors (Oppenheim 1991; Davies 2003). This model, called neurotrophic theory, has since expanded to take into account competition for afferent inputs (Linden 1994; Sendtner et al. 2000; Banks et al. 2005). Hence, cellular survival would be regulated by competition for retrograde and anterograde trophic factor delivery that depends on electrical activity from the afferents and target neurons. PCD during early development would serve to match the size of interconnected neuronal populations (Galli-Resta and Resta 1992; Posada and Clarke 1999; Buss et al. 2006). As synaptogenesis and refinement of the cortical network precedes (Innocenti and Price 2005), the concomitant neuronal loss could correspond to competition for survival as conceptualized in the neurotrophic theory. In such a model, neurons that integrate into the developing cortical network would develop stable electrical activity and would then be reinforced by trophic factors supply, whereas neurons that would fail to incorporate into the network would lack activity and would be eliminated by entering PCD. Support for the hypothesis of activity-dependent survival comes from in vitro and in vivo studies on cortical neurons that have demonstrated the dependence of survival on the activation of the N-methyl-D-aspartic acid receptor (NMDAR) and subsequent brain-derived neurotrophic factor (BDNF) synthesis/release (Ghosh et al. 1994; Ikonomidou et al. 1999; Hardingham et al. 2002; Yoon et al. 2003; Hansen et al. 2004; Walz et al. 2006). Several questions, however, are not yet answered. Whether PCD in the postnatal cerebral cortex in vivo targets neurons has not been unequivocally clarified. Furthermore, the patterns of neuronal activity that drive the survival of early postnatal cortical neurons are largely unknown. It is unclear whether the factors for neuronal survival defined in dissociated cell cultures are also applicable in more intact networks. We aimed at correlating neuronal activity and apoptosis in neurons from the developing network of the cerebral cortex. For this purpose, we have used organotypic slice cultures of somatosensory cortex from newborn mice. By combining measurements of apoptosis and recording of neuronal activity, we have analyzed which activity patterns could promote neuronal survival.

Keywords: apoptosis, cell death, development, neocortex, NMDA, spontaneous activity

Introduction Programmed cell death (PCD) is an essential process for the proper development of the nervous system. In the developing cerebral cortex, up to 25% of the neuronal population is eliminated during early postnatal stages as revealed by histological analysis (Heumann et al. 1978; Heumann and Leuba 1983; Ferrer et al. 1992; Voyvodic 1996) and labeling of apoptotic DNA fragmentation (Spreafico et al. 1995; Verney et al. 2000). The decision to enter the PCD process depends on several factors such as the proliferative or metabolic state of the neuron as well as the presence of different extracellular signals (Hidalgo and Ffrench-Constant 2003; Benn and Woolf 2004; Buss et al. 2006). However, the question which initial factor drives cortical neurons into apoptosis is still unresolved. Based on the knowledge from the motor and sympathetic nervous systems, it has been proposed as a general developmental rule that an exuberant number of neurons is produced, from which a subset is selected according to the ability to connect to the appropriate target, that delivers Ó The Author 2007. Published by Oxford University Press. All rights reserved. For permissions, please e-mail: [email protected]

1

Institute of Physiology and Pathophysiology, University of Mainz, Duesbergweg 6, D-55128 Mainz, Germany and 2 Current address: Volkmar Lessmann Institute of Physiology, Otto-von-Guericke-University, 39120 Magdeburg, Germany

Material and Methods In Vivo Analysis of Apoptosis All experiments were conducted in accordance with the national laws for the use of animals in research (Deutsche Forschungsgemeinschaft 2004) and approved by the local ethics committee. Postnatal C56/BalbC mice (P3, P6, and P10) were deeply anaesthetized with an intraperitoneal injection (800 ll) of pentobarbital and perfused intracardially with 4% paraformaldehyde (PFA) in 0.1 M phosphate buffer, pH 7.4 at 4 °C. The brain was removed from the skull and placed

for 2 h at 4 °C into 4% PFA. After dehydration in graded ethanol baths (3 times for 15 min in 50%, 70%, and 95% and 3 times for 1 h in 100% ethanol), brains were incubated in 100% butanol for 12 h and then impregnated at 60 °C with 50% butanol/50% paraffin (Rotiplast, Roth, Karlsruhe, Germany) for 48 h before final incubation in 100% paraplast for 24 h. Paraffin sections of 10 lm thickness were cut on a microtome (Supercut 2050; Leica, Wetzlar, Germany). The sections were treated with xylol and butanol and rehydrated in ethanol baths (100, 95, 70, 50, and 0%) with a final step in 0.01 phosphate buffer saline (PBS). To improve the efficiency of the Terminal deoxynucleotidyl Transferase Biotin-dUTP Nick End Labeling (TUNEL), sections were incubated in 0.01 M citrate buffer pH 3 and heated 3 times for 5 min in the microwave at 350 watts. The sections were washed in PBS and incubated with 10% terminal deoxynucleotidyl transferase solution in labeling solution containing deoxyuridine coupled with fluorescein (Roche Applied Science, Mannheim, Germany) for 60 min at 37 °C in the dark. The signal was amplified by incubating the slices either with cy2-coupled antibody directed against fluorescein (1/400; Jackson Immunoresearch Laboratories Inc., West Grove, PA) or with a peroxidasecoupled antibody directed against fluorescein (Roche). In the latter case, the signal was obtained by peroxydase reaction with 0.3% w/v 3,3#diaminobenzidine (DAKO, Carpinteria, CA) in Tris buffer with 0.3% v/v H2O2. For cresyl violet staining, the following procedure was used: 5 min in 0.1% cresyl violet (Merck, Darmstadt, Germany), 5 min in water, 2 min in 70% ethanol, 2 min in 96% ethanol, and 2 min in 100% ethanol. For costaining of TUNEL and caspase-3, the treatment with citrate buffer was performed at room temperature. Following TUNEL reaction, the sections were incubated overnight with the primary antibody specifically directed against the activated form of caspase-3 (a-casp3) (asp175, 1/400; Cell Signaling Technology Inc., Pharmingen, San Diego, CA). After washing with PBS, sections were incubated with biotinylated secondary antibody directed against rabbit (1/400; Vector Laboratories, Burlingame, CA) and cy2 coupled antibody directed against fluorescein (1/400; Jackson Immunoresearch) for 2 h. Finally, sections were incubated with streptavidin coupled to cy3 (1/800; Jackson Immunoresearch). For immunostaining for activated caspase-3 not coupled with TUNEL, 100 lm vibratome sections were cut. After permeabilization with 0.1% Triton X-100 for 1 h, immunostaining was performed as described above. Preparation of Organotypic Neocortical Slice Cultures Organotypic neocortical slices were prepared from newborn (P0) C56/ balbC mice and cultivated according to the Stoppini method (Stoppini et al. 1991). Previous studies have demonstrated that organotypic slice cultures can be used effectively to investigate the molecular and cellular mechanisms underlying apoptosis during early cortical development (Haydar et al. 1999). Animals were killed by decapitation and the brain was quickly removed. All subsequent procedures were performed in Minimal Essential Medium (MEM, Gibco, Invitrogen, Carlsbad, CA) supplemented with 2 mM glutamine, pH 7.4 at 4 °C. Neocortical hemispheres were isolated from the hippocampus, thalamus and striatum and 350 lm thick coronal slices containing the parietal cortex were cut with a tissue chopper (McIlwain, Mickle Laboratory Engineering, Surrey, United Kingdom). The meninges were carefully removed and the dorsolateral part of the cerebral cortex was dissected with a microscalpel. The isolated parietal cortical slices were transferred onto Millicell-CM membranes (Millipore, Bedford, MA) placed into 35 mm petri dish. Slices were kept at 37 °C, 5% CO2 in a serum-free medium containing neurobasal medium supplemented with 2% B27 (Gibco), 2 mM glutamax (Gibco) and 10 lg/ml penicilline/streptomycine. After 24 h, 1 lM Arabinofuranoside (AraC, Sigma-Aldrich, Steinheim, Germany) was added to the medium. Culture medium with AraC was renewed every two days, thereafter. Pharmacology After 4 to 5 days in culture, the medium was replaced by neurobasal medium without B27 and specific antagonists were added. In controls an equal volume of the antagonist solvent was applied. AraC was omitted during pharmacological treatments of the slices. After incubation periods of 6, 12 or 24 h, slices were fixed for 30 min with 4% PFA.

1336 Apoptosis in Developing Cortex

d

Heck et al.

The following drugs were used: 1 lM tetrodotoxin (TTX) citrate (Tocris Cookson, Bristol, UK); 10 lM 6-cyano-7-nitroquinoxaline-2,3dione (CNQX, Sigma); 80 lM 1-(4-Aminophenyl)-4-methyl-7,8-methylenedioxy-5H-2,3-benzodiazepine (GYKI 52466, a gift from Dr I. Vila´gi, Eo¨tvo¨s Lora´nd University, Budapest, Hungary); 20 lM (±) 3-(2carboxypiperazin-4-yl)propyl-1-phosphonic acid (CPP, Sigma); 400 nM [(R)-[(S)-1-(4-bromo-phenyl)-ethylamino]-(2,3-dioxo-1,2,3,4-tetrahydroquinoxalin-5-yl)-methyl]-phosphonic acid (NVP-AMM077, Novartis Pharmaceuticals, Basel, Switzerland); 3 lM ifenprodil (Tocris); 10 lM nifedipine (Sigma); 100 lM gabazine (SR-95531, Sigma) 25 lM Mefloquine (Roche); 200 nM k252a (Alomone labs, Jerusalem, Israel). Drugs were stored at –20 °C and diluted to the final concentration on the day of experiment. Immunostaining for Activated Caspase-3 and Quantification of Apoptotic Cell Death An antibody directed against the cleaved, active form of caspase-3 was used (Asp175, Cell Signaling Technology Inc.) (Marin-Teva et al. 2004). This antibody is specific for the activated caspase-3 because it does exclusively recognize the 17/19-kDa fragment of the cleaved caspase-3 form (Lavoie et al. 2005). After fixation slices were rinsed in PBS, treated for 30 min with 0.1% Triton X-100 in PBS, and preincubated in 5% bovine serum albumin (BSA) in PBS for 2 h. Primary rabbit polyclonal antibody against activated caspase-3 (asp175, 1/400, Cell Signaling Technology Inc.) was incubated overnight at room temperature in 5% BSA in PBS. Sections were washed in PBS and incubated with the secondary antibody (biotin-conjugated donkey polyclonal antibody directed against rabbit 1/400, Vector Laboratories) for 2 h at 4 °C, then washed in PBS, and finally incubated with alexa488-conjugated streptavidin (Molecular Probes, Eugene, OR). After washing with PBS, sections were mounted on slides. For double staining with glial fibrillar acidic protein (GFAP), the slices were immunostained for a-casp3 as described above and then fixed again with 4% PFA for 15 min. Immunostaining with rabbit polyclonal antibody directed against GFAP (1/400, Dako, A/S, Denmark) was then processed using the same protocol as for a-casp3. Slices were analyzed using a confocal system (QLC100; Visitech, Sunderland, UK) attached to a Olympus microscope (BX 51 WI; Olympus, Tokyo, Japan). Immunostained whole-cortex slices were observed at 6003 magnification at an upright microscope (BX51WI) which corresponded to a field of view of 125 lm2. Apoptotic cells were counted in 18 fields of view that were assigned to cover all cortical layers in 3 different regions of each slice. Because there were no significant differences between the regions and in between the layers, the number of apoptotic cells from the 18 fields of view was averaged. For each field of view, the focus was manipulated to count cells in the whole depth of the slice. Permeabilization of the slice prior to immunostaining allowed the observation of positive cells in the depth of the slice. Using this standardized protocol, apoptotic cells were manually counted by 3 independent observers naive to the experimental condition (in a blind manner). Slices from at least 3 independent cultures were analyzed. The results are shown as percentage of the respective control condition (control is 100%).

Electrophysiological Recordings All electrophysiological recordings were performed in an artificial cerebrospinal fluid (ACSF) that resembled the ionic and molecular composition of the neurobasal medium without B27 (Brewer et al. 1993). Electrophysiological recordings were performed in organotypic neocortical slices after 4--5 days in culture both at the network level using a multielectrode array (MEA) and at the cellular level using whole-cell patch-clamp recordings. Spontaneous activity was recorded in control conditions, and pharmacological agents were added after at least 4 spontaneous events had been observed. Cultured slices were transferred to 3-D multi-MEA chips (interelectrode distances of 200 lm, electrode diameter of 30 or 40 lm; Ayanda Biosystems, Lausanne, Switzerland) on MEA 1060-INV-BC (Multi Channel Systems [MCS], Reutlingen, Germany), which were mounted on an inverted microscope (Optika microscopes, Ponteranica, Italy). Slices were superfused with an ACSF consisted of (in millimolar) 51.3

NaCl, 26 NaHCO3, 0.9 NaH2PO4 H2O, 0.812 MgCl2 6H2O, 1.8 CaCl2, 5.36 KCl, 25 D-glucose, 10 (N-[2-hydroxyethyl]-piperazine-N#-[2-ethanesulfonic acid]) sodium salt, 0.23 sodium pyruvate, 0.5 L-glutamine, 0.4 glycine, 0.4 L-serine, 0.02 L-alanine, 0.028 choline chloride, and 0.2 lM Fe(NO3)3 H2O equilibrated with 95% O2/5% CO2 (pH 7.4; osmolarity 205 mOsm). ACSF was continuously perfused at a rate of 2--3 ml/min at a temperature of 28 °C. Field potentials were recorded simultaneously with 60 extracellular electrodes at a sampling rate of 1 kHz using the MEA_RACK software (MCS). Data were imported to a custom-written program in Matlab version 6.5 (Mathworks, Natick, MA) with datastrm.m and nextdata.m (MC_Rack, MCS). Field potentials in 60 channels were plotted, and the duration (interval between onset and end of one complete oscillatory cycle) and amplitude (voltage difference between the maximal positive and negative peak) of oscillatory field potential activity was determined. Time-frequency plots were calculated by transforming field potentials using Morlet continuous wavelet (Matlab, cwt.m command) and power values were normalized according to the formula (power values – minimal)/(maximal – minimal). Relative power (0--1) was displayed in blue and red, respectively. To perform whole-cell patch-clamp recordings from single neurons, slices were transferred to a submerged-type recording chamber (Luigs and Neumann, Ratingen, Germany) mounted on the fixed stage of an upright microscope (BX51WI, Olympus). The ACSF flow rate was 3 ml/ min, and the temperature was maintained at 28 °C by the use of a temperature control unit (Luigs and Neumann). Single neurons were visualized by infrared differential interference contrast with a video system allowing contrast enhancement (VX55, Till Photonics, Grafelfing, Germany). Patch pipettes were pulled from borosilicate tubing (GB200F-8P with filament, 2.0 mm outside diameter, 1.16 mm inside diameter; Science Products, Hofheim, Germany) on a vertical puller (PP-830; Narishige, Tokyo, Japan). The pipettes were filled with an intracellular solution containing (in millimolar) 82 K-gluconate, 8 KCl, 1 CaCl2, 11 ethyleneglycol-bis(2-aminoethylether)-N, N, N#, N#tetraacetic acid, 10 (N-[2-hydroxyethyl]-piperazine-N#-[2-ethanesulfonic acid]) potassium salt, 2 Mg2-adenosine triphosphate and 0.5 Naguanosine triphosphate (pH adjusted to 7.4 with KOH, osmolarity measured 225 mOsm). In all experiments, 0.5% biocytin (Sigma) was added to the pipette solution to label the recorded cells. When filled with intracellular solution, the pipettes had a resistance ranging from 15 to 25 MOhm. Signals were directly amplified, low-passed filtered at 2.9 kHz, and digitized by EPC10 amplifier (HEKA, Lambrecht, Germany). Data were recorded and analyzed with WinTIDA 5.17 software (HEKA). The resting membrane potential was measured immediately after obtaining the whole-cell configuration. The passive electrophysiological properties of the cells and their firing patterns were characterized under current-clamp condition. Current voltage relationships were assessed by recording responses to series of 2-s long hyper- and depolarizing current pulses from a holding potential of –60 mV. Spike amplitude was measured from action potential threshold to the peak, and the spike duration was determined at the half-maximum spike amplitude. Histology and Morphological Analysis of the Recorded Cells All recorded cells were filled with 0.5% biocytin (Sigma). After recording, slices were fixed in 4% buffered PFA solution for at least 24 h, rinsed in PBS, and incubated for 2 h in alexa488 coupled to streptavidin (1/800, Molecular Probes). Slices were analyzed using a confocal system (QLC100, Visitech) attached to an Olympus microscope (BX 51 WI). Images were collected with a 203 objective (Olympus) and captured using a cooled charged couple device camera (CoolSnap HQ; Roper Scientific, Trenton, NY). Because the field of view was too narrow to cover a whole neuron, mosaic images were assembled.

Measure of Levels of Phosphorylated CREB For the measurment of the levels of the phosphorylated form of cAMP response element--binding protein (CREB), slices were treated for 6 h + with CPP, high [K ]e, and k252a as described in the Results and then harvested in a lysis buffer composed of 0.5% Triton X-100, 1 mM EGTA, 6 M Urea in PBS. The slices were allowed to sit on ice for 15 min and

were briefly vortexed before use. Enzyme-Linked ImmunoSorbent Assay (ELISA) assay was performed according to manufacturer’s protocol (DuoSet IC Phospho-CREB S133, R&D systems, Minneapolis, MN). Briefly, capture antibody directed against phospho-CREB in PBS was incubated overnight in maxisorp 96-wells plates (Nunc, Wiesbaden, Germany). After washing in PBS and blocking step in 1% BSA in PBS, samples were incubated for 2 h followed by incubation for 2 h with detection antibody directed against phospho-CREB in 1% BSA in PBS. After washing step, 100 lL streptavidin coupled with horseradish peroxydase was incubated for 20 min. After several washes in PBS, the peroxydase activity was detected by addition of 10% w/v 2,2#-azino-bis(3-ethylbenzthiazoline-6-sulfonic acid) (ABTS, Sigma) diluted with 0.025% v/v water peroxide 30% in 0.05 M phosphate-citrate buffer. The reaction was allowed to develop overnight at 4 °C and was measured at 405 nm on an ELISA plate reader (Infinite 200; Tecan, Crailsheim, Germany). Statistics For statistical analyses, Student’s t-test and analysis of variance (ANOVA) followed by Newman--Keuls post hoc test were used. The nonparametric Mann--Whitney and Kruskal--Wallis test (nonparametric equivalent of ANOVA) followed by Dunn post hoc test were used when the sample were not following a Gaussian distribution. All data are expressed as mean ± standard error of the mean. Number of apoptotic cells are represented in histograms as percentage of control values, the control being normalized at the value 100%.

Results Caspase-3 Dependent Apoptosis In Vivo In order to address the question whether PCD during early cortical development in vivo involves the activation of caspase3, we performed TUNEL and caspase-3 staining in the somatosensory cortex of postnatal day mice. In agreement with previous studies (Verney et al. 2000), scattered TUNELpositive cells could be observed in all cortical layers (Fig. 1 A,B). The occurrence of caspase-3--dependent PCD was assessed by immunostaining directed against the a-casp3. A-casp3--positive cells could be detected in all cortical layers (Fig. 1C,D). The number and distribution of TUNEL stained and a-casp3--positive cells were very similar, suggesting that both methods label the same population of cells. This hypothesis was supported by our double staining experiments, which demonstrated a colocalization of TUNEL and a-casp3 signals (Fig. 1E). Because activation of caspase-3 is an early event in PCD (Marin-Teva et al. 2004), a-casp3--positive cells still showed a relatively normal morphology. The detailed microscopical analysis of a-casp3--positive cells revealed a clear neuronal morphology (Fig. 1F). Pyramidal neurons with typical soma morphology and proximal dendrites could be identified in all cortical layers. However, in some pyramidal neurons, primary dendrites were interrupted and the soma showed a more round shape, indicating that these neurons were already affected by the cell death process. Our data clearly indicate that neurons in all cortical layers of the newborn mouse cerebral cortex undergo caspase-3-dependent apoptosis in vivo. Caspase-3--Dependent Apoptosis In Vitro In order to study the role of electrical activity and the contribution of the different glutamate receptors in the apoptosis process during early cortical development, we established organotypic cultures of newborn mouse somatosensory cortex (Fig. 2A) and modified the network activity by pharmacological interventions. Biocytin-filled pyramidal Cerebral Cortex June 2008, V 18 N 6 1337

Figure 1. Neurons undergo caspase-3--dependent cell death in the neonatal mouse cerebral cortex in vivo. (A--C) Analysis of cell death in the P3 mouse somatosensory cortex in vivo. (A) Nissl staining showing the 6-layered organization of the neonatal mouse cerebral cortex. (B) TUNEL staining, which labels DNA fragmented cells, demonstrates PCD in the neonatal cortex (arrows). (C) Staining for activated caspase-3 in the P3 mouse somatosensory cortex (arrows). The distribution and number of positive cells are similar to that demonstrated with the TUNEL method. (D) Cells positive for the activated caspase-3 are similarly found in the P6 mouse somatosensory cortex in vivo. (E) Double staining for TUNEL (signal in green localized in the nucleus) and activated caspase-3 (signal in red localized in the soma) in the P10 mouse cerebral cortex in vivo. Note that double labeled cells show already structural alterations and fragmentation of the nucleus. (F) Staining for the activated caspase-3 in P3 (upper row) and P6 neocortex (lower row) in vivo. Confocal pictures allow neuronal identification, some cells resemble pyramidal neurons, in which the cell death process can be identified by a loss of morphology with interrupted dendrites (arrows) and round soma (arrowhead). Scale bar in (D) corresponds to 200 lm in (A--D); scale bars in (E--F) 10 lm. All photographs are from somatosensory cortex.

neurons in our neocortical slice cultures revealed the typical complex morphology with a rich dendritic branching pattern and axons extending over several layers up to 850 lm in length (Fig. 2A,B). To study caspase-3 dependent apoptosis in vitro, organotypic neocortical slices were immunostained with an antibody directed against a-casp3 (see Fig. 2D). During the first 2 days in vitro (div), slice cultures exhibited a pronounced a-casp3 immunoreactivity. At 3 div, the number of a-casp3--positive cells markedly decreased and remained constant until 12 div, the latest developmental stage analyzed in the present study. All subsequent experiments were performed on slice cultures between 4 and 6 div because during this period the apoptosis rate had already reached the plateau phase and because neocortical slice cultures at this stage reveal spontaneous activity patterns similar to those observed in the early postnatal rodent cortex in vivo (see below and Khazipov and Luhmann 2006). We first studied the dependence of the slice cultures on external supply of trophic factors by replacing the culture medium, which was composed of neurobasal medium supplemented with B27, with neurobasal medium without B27. After incubation periods of 6, 12, or 24 h, slice cultures were immunostained for a-casp3 and the number of positive cells was determined. The number of a-casp3--positive neurons remained largely constant over 24 h indicating that this deprivation from the B27 medium supplement did not induce any increase in the number of apoptotic cells (Fig. 2C). In contrast, a previous report in dissociated cortical cultures (Poser et al. 2003; Liot et al. 2004) and our own observations on an increased cell death in dissociated cultures of cortical neurons deprived of B27 medium supplement (A. Golbs and H.J. Luhmann, unpublished data) demonstrate a difference between both culture systems in terms of sensitivity to trophic factors external supply. These observations indicate that 1338 Apoptosis in Developing Cortex

d

Heck et al.

neurons in organotypic cultures develop within an environment that allows for the release and constrains the diffusion of trophic factors supporting neuronal survival. Activity-Dependent Control of Neuronal Apoptosis To assess whether survival is dependent on electrical activity, slice cultures were treated with 1 lM tetrodotoxin, a blocker of voltage-dependent sodium channels (Fig. 2E). The number of apoptotic cells was significantly (P < 0.001, n = 12 slices for each time point) higher after 6, 12, and 24 h of TTX application as compared with the control slices without TTX (Fig. 2F). This increase in apoptotic cells could be observed uniformly in the whole slice culture, indicating that all cortical layers showed the same sensitivity to neuronal activity. The possibility of cell death in astrocytes was assessed by double staining for a-casp3 and the marker GFAP. In these analyses, we found no evidence for colocalization of a-casp3 and GFAP, suggesting that TTX induced neuronal apoptosis and not glial cell death (Fig. 2G--I). These TTX experiments were performed in the absence of B27 supplement. In culture medium supplemented with B27, TTX application did not induce any significant changes in the apoptosis rate (data not shown), indicating that B27 can compensate for the lack of trophic factors released by neuronal activity. For this reason, all subsequent experiments were performed in neurobasal medium in the absence of B27. Electrophysiological and Pharmacological Characterization of Spontaneous Network Activity in Neocortical Slice Cultures Because TTX treatment caused an increase in the apoptosis level, we expected that spontaneous electrical activity patterns sustain neuronal survival. Therefore, we studied the spontaneous network activity in neocortical slice cultures at the network and single-cell level by the use of different electrophysiological recording techniques. The slice culture was

Figure 2. Intact cortical organization and cellular morphology in organotypic neocortical slice cultures and sensitivity to B27 deprivation. (A) Photograph of a 5 div neocortical slice culture illustrating the overall morphology of the slice. The picture of a biocytin-filled pyramidal neuron has been inserted to illustrate that neurons extend ramified axons throughout all layers. (B) Examples of biocytin-filled pyramidal neurons in slice cultures at 5 div. The dendrites are well developed and reveal numerous branches. (C) Serum deprivation does not cause enhanced cell death in slice culture. Compared with cultures maintained for 6 h in neurobasal medium supplemented with B27, removal of B27 for 6 h does not induce a significant change in the number of a-casp3--positive cells (Mann--Whitney test, P [ 0.05). The level of apoptosis remains constant for 12 and 24 h (Kruskal-Wallis test, P [ 0.05). Data show the number of cells stained for activated caspase-3. (D) Pattern of a-casp3 immunoreactivity in a 5 div organotypic neocortical slice in control condition. The apoptotic cells are homogenously distributed throughout all neocortical layers. (E) In comparison to control condition, slices treated with 1 lM tetrodotoxin show a prominent increase in the number of cells stained for the a-casp3. Insets in (D) and (E) show a-casp3--positive cells at higher magnification. Note the typical neuronal morphology. (F) Counting of caspase-3--positive cells reveals a significant increase in the number of apoptotic cells in TTX-treated slices as compared with the controls (5100%). After only 6 h of TTX treatment, the amount of apoptotic cells reached 250% of the control (Mann--Whitney test, P \ 0.0001). This effect is maintained for 12 h (Mann--Whitney test, P \ 0.0001) and 24 h (P \ 0.0001) of TTX treatment. (G--I) Cell death of astrocytes was not detected in TTX-treated slices. A-casp3 was detected in cells with neuronal morphology (G) but did not colocalize with the astrocytic marker GFAP (H, I). Scale bar in (A), (D) and (E) 5 200 lm, in (B) 5 50 lm, inset in (D) and (E) 5 10 lm. Scale bar in (G--I) 5 20 lm.

placed onto a MEA system (Fig. 3A), and the spontaneous neuronal activity was recorded with 60 extracellular electrodes with the MEA system. In agreement with previous in vivo and in vitro observations in newborn rodent cerebral cortex (Kilb and Luhmann 2003; Khazipov et al. 2004; Dupont et al. 2006; Hanganu et al. 2006), neocortical slice cultures showed synchronized network oscillations. Spontaneous network oscillations could be recorded simultaneously over the whole MEA and in all cortical layers (Fig. 3B). The spontaneous network oscillations occurred at an average interval of 6.7 ± 0.8 min (n = 8 slices) had a duration of 9.1 ± 0.65 s and a maximal amplitude of 679 ± 18 lV (n = 8 slices). Each event was characterized by an initial oscillation in the frequency range from 5 to 8 Hz, followed by repetitive wave-like discharges (Fig. 3C). In 5 slices, the activity appeared simultaneously at all recording sites, and in 9 slices, the network oscillations propagated with an average velocity of 70.3 ± 4.9 mm/s over the MEA. Thus, the oscillatory activity in our neocortical slice cultures was highly correlated over a few hundred micrometers (Fig. 3D,E). The pattern and pharmacology of the spontaneous network activity in organotypic neocortical cultures was also studied at the cellular level by performing whole-cell patch-clamp recordings from visually identified and biocytin-stained pyra-

midal neurons (see Fig. 2B). All recorded neurons (n = 40 neurons) exhibited relatively uniform membrane properties (Fig. 4A-C). The average resting membrane potential was –49.5 ± 1.5 mV and the input resistance amounted to 0.78 ± 0.12 GOhm (n = 34 neurons). All neurons were capable of firing repetitive action potentials in response to sustained depolarization (Fig. 4A). Action potential threshold was found to be at –29.5 ± 1.5 mV. Average action potential amplitude and half-maximal durations were 39.7 ± 1 mV and 7.4 ± 0.5 ms (n = 36), respectively (compare Fig. 4B). Current-voltage relationship showed no rectification (Fig. 4C). Taken together, these properties correspond to recordings of immature pyramidal neocortical neurons, as determined in acute slice preparations (Luhmann et al. 2000). In 20 out of 28 neurons recorded for at least 20 min, spontaneous burst discharges with an average spike frequency of 7.4 ± 0.5 Hz and duration of 5.8 ± 0.9 s could be observed at intervals of 6.7 ± 0.7 min (Fig. 4D). The properties of these cellular events resembled those of the extracellular network bursts recorded with the MEA. In the remaining 8 neurons, we did not observe any spontaneous activity within the observation period of at least 20 min. Spontaneous synaptic events were also recorded in voltage-clamp mode (Fig. 4E). Cerebral Cortex June 2008, V 18 N 6 1339

Figure 3. Synchronized spontaneous network activity in neocortical slice cultures. (A) Photograph of a 7 div organotypic neocortical slice culture placed on a 60-channel MEA. Black dots represent extracellular recording electrodes, separated by 200 lm. (B) Simultaneous 60-channel MEA recording from neocortical slice culture shown in (A). Four channels are damaged and were switched off. (C) Spontaneous oscillatory activity recorded with channel 45 and corresponding wavelet analysis. (D) Color plot illustrating the correlation coefficient calculated from cross-correlation analyses between channel 45 and all other channels. (E) Delay of activity onset recorded with channel 45 as compared with the activity onset in the other channels.

Spontaneous synaptic currents occurred in bursts with an average duration of 7.09 ± 0.48 s (n = 19 slices). The first synaptic current always showed a much higher maximal amplitude with an average of 65.5 ± 4.6 pA and a half-maximum duration of 333 ± 35 ms (n = 15). Both, the postsynaptic potentials as well as the postsynaptic currents consisted of multiple components, indicating that this spontaneous activity was generated by a polysynaptic network. Finally, we studied with the patch-clamp technique the effect of TTX, which caused a prominent increase in neuronal apoptosis (see Fig. 2D--F), on the spontaneous activity at the single-cell level. In all investigated neurons (n = 3) spontaneous burst discharges were completely blocked by TTX (Fig. 4F).

a significant increase in the number of apoptotic cells (P < 0.001, n = 9 slices for each time point, Fig. 5F). Highmagnification microscopic analysis of the a-casp3--positive cell confirmed the neuronal nature of the apoptotic cells (Fig. 5E). Next, we studied by whole-cell voltage-clamp recordings the role of the different ionotropic glutamate receptors in generating the spontaneous synaptic activity. Spontaneous postsynaptic burst discharges recorded in neurons from 5 div slice cultures were completely blocked by CNQX (n = 4 slices, Fig. 5G) and CPP (n = 4 slices, Fig. 5H), demonstrating the functional role of AMPA/kainate receptors and NMDA-Rs, respectively, in generating spontaneous network activity in neocortical slice cultures.

Role of Ionotropic Glutamate Receptors The putative role of AMPA/kainate-type glutamate receptors was studied by incubating slice cultures in 10 lM of the antagonist CNQX (Fig. 5A). After 6, 12, or 24 h of incubation, CNQX-treated slices showed a significantly (P < 0.001, n = 9 slices for each time point) higher number of apoptotic cells as compared with the untreated controls. The specific role of AMPA receptors was further assessed by application of the AMPA receptor antagonist GYKI 52466 (Fig. 5B). For all 3 incubation times, 80 lM GYKI induced a highly significant (P < 0.001, n = 6 slices for each time point) increase in the apoptosis rate as compared with untreated controls demonstrating that the activation of AMPA glutamate receptors is essential for neuronal survival. The role of NMDA-R in the regulation of activity-dependent apoptosis during early cortical development was studied by the use of the NMDA-R antagonist CPP (Fig. 5C-F). For all 3 incubation times blockade of NMDA-R caused

Activation of L-type Ca+ Channels as a Secondary Survival Pathway Our results suggested that calcium influx via NMDA-Rs is a critical determinant for neuronal survival in our slice cultures. To address the question whether other sources of calcium influx can substitute for the NMDA-R--dependent survival response, cells were depolarized by elevating of the extracellular potassium concentration ([K+]e) to 25 mM, thus activating voltage-gated calcium channels. When the CPP-treated slices were incubated in culture medium containing high [K+]e, the number of apoptotic cells was reduced to control values observed in the absence of CPP and low [K+]e (Fig. 6D). These data indicate that high [K+]e induces an alternative survival pathway that can overcome the blockade of NMDA-R. + To understand the activity pattern in CPP and high [K ]e that promotes survival, MEA multichannel recordings were performed from slices cultures treated with 10 lM CPP and 25 mM

1340 Apoptosis in Developing Cortex

d

Heck et al.

Figure 4. Electrophysiological properties of pyramidal neurons in organotypic neocortical slice cultures. (A) Voltage responses to injection of hyperpolarizing and depolarizing current pulses in a 6 div neuron. Suprathreshold activation elicits repetitive single action potentials. (B) First action potential of the trace illustrated in (A) (B1) and first derivative of voltage (B2). Note relatively small amplitude and slow time course of the action potential, which are characteristic features of immature neocortical pyramidal neurons. (C) Current--voltage relationship for the recordings shown in (A). (D) Spontaneous burst discharge recorded in current-clamp mode from a 5 div neuron. (E) Two typical examples of spontaneous events recorded in voltage-clamp mode at 60 mV in a 5 div neocortical slice culture. The first trace is from the same neuron as in (D). (F) Spontaneous activity was completely blocked by the addition of 1 lM TTX.

[K+]e (Fig. 6). In agreement with our single cell recordings, spontaneous events were completely blocked by CPP (n = 5 slices, Fig. 6B). Upon addition of 25 mM extracellular potassium to the CPP-containing bathing solution, a pronounced increase in activity with a maximal amplitude of 332 ± 97 lV could be observed (n = 3 slices, Fig. 6C). However, this network activity with an average duration of 51.6 ± 16.2 s was observed only transiently during the washing phase of 25 mM potassium. Thereafter, no further network activity could be recorded for observation periods up to 1 h. It has been demonstrated previously in neonatal mouse cerebral cortex that spontaneous, synchronous electrical activity could be blocked by the antagonist of L-type calcium channels nifedipine (Corlew et al. 2004). The functional role of these voltage-dependent calcium channel on the network activity in neocortical slices cultures was studied by application of 10 lM nifedipine. Under control conditions in normal extracellular solution, nifedipine had no significant effect on the apoptosis rate (Fig. 6D).

However, in cortical slice cultures treated with 25 mM extracellular potassium and CPP, nifedipine caused a significant (P < 0.01, n = 4 slices) increase in the number of apoptotic cells, demonstrating that under blockade of NMDA-R and increased neuronal activity, L-type calcium channels have a prosurvival influence (Fig. 6D). We then addressed the question, whether the high [K+]e induced increase in electrical activity could rescue the neurons via the release of trophic factors. Slices were treated with 200 nM k252a, an inhibitor of tyrosine kinases of the neurotrophin receptors TrkB and TrkC. In comparison to CPP-treated slices, the addition of 25 mM of [K+]e rescued the neurons, but the prosurvival effect of high [K+]e was prevented in the presence of 200 nM k252a (Fig. 6E). Hence, the high [K+]e-induced rescue effect is dependent on the activation of tyrosine kinases of the neurotrophin receptors TrkB or TrkC. It has been previously shown in dissociated cell cultures that high [K+]e activates survival signaling pathways that involve the Cerebral Cortex June 2008, V 18 N 6 1341

Figure 5. Activation of ionotropic glutamate receptors is essential for neuronal survival. (A) Blockade of AMPA/kainate glutamate receptors with 10 lM CNQX induces increase in the number of apoptotic cells compared with control condition. This increase is observed after 6, 12, and 24 h of treatment with CNQX (Mann--Whitney test, all P \ 0.0001). (B) The specific AMPA receptor blocker GYKI (80 lM) also induces an increase in apoptosis rate after 6, 12, and 24 h of treatment (Mann--Whitney test, all P \ 0.001). (C, D, E) In comparison to untreated control slices (C), slices treated with the NMDA receptor antagonist CPP show a homogeneous and prominent increase in the number of apoptotic cells stained for a-casp3 (D). Higher magnification of a-casp3--positive cells allows the identification of neurons with a typical somatodendritic morphology (E). As indicated by arrows, some neurons show relatively intact morphology (arrow), whereas other neurons reveal typical signs of apoptosis with round soma and degenerating processes (arrowhead). (F) Quantification of apoptosis rate by counting a-casp3--positive cells after NMDA receptor blockade reveals significant increase after 6, 12, and 24 h of treatment with 10 lM CPP (Mann--Whitney test, all P \ 0.0001). Scale bar in (C) and (D) 5 200 lm; in (E) 5 10 lm. (G--H) Spontaneous activity recorded at single-cell level by patchclamp technique in voltage-clamp mode was completely blocked by the addition of either 10 lM CNQX (G) or 10 lM CPP (H).

phosphorylation of CREB (Dolmetsch et al. 2001). We measured in our organotypic neocortical slice cultures, the level of CREB phosphorylation at the Serin 133 site (P-CREB). The addition of high [K+]e increased P-CREB levels, and this increase was not observed in the presence of 10 lM nifedipine (Fig. 6F). Hence, the activation of L-type calcium channels by high [K+]e induces the activation of the factor CREB, which has been shown to be involved in pathways promoting neuronal survival (Hardingham and Bading 2003). The NMDA Receptor Subunits NR2A and NR2B Have Different Effects on Neuronal Survival To elucidate the contribution of different NMDA-R subunits in the NMDA-R-dependent apoptosis process, we blocked the NR2A subunit with 400 nM NVP-AMM077 (Berberich et al. 2005) and the NR2B subunit with 3 lM ifenprodil (Williams 1993). The NR2A subunit antagonist NVP-AMM077 induced a complete blockade of the spontaneous network activity recorded in normal extracellular solution (n = 3 slices, Fig. 7A). In accordance with these electrophysiological results, blockade 1342 Apoptosis in Developing Cortex

d

Heck et al.

of the NR2A subunit also induced a significant (P < 0.001, n = 12 slices for each time point) increase in the apoptosis rate after 6, 12, and 24 h when compared with the untreated controls (Fig. 7B). Interestingly, the number of apoptotic cells was highest after 6 h treatment with NVP-AMM077 and significantly decreased with longer application times (P < 0.001 in between each time point), indicating that a NR2Aindependent process is upregulated after 12--24 h. Indeed, synchronous activity was observed in slices (n = 4) preincubated for 24 h in NVP-AMM077, which indicates that changes in the mechanisms that generate spontaneous activity occur to compensate for the blockade of NR2A-containing NMDA-R (data not shown). The NR2B subunit specific antagonist ifenprodil had no significant effect on the spontaneous network activity (Fig. 7C) and on the apoptosis rate after 6 h application time (P > 0.05, n = 9 slices), but the number of apoptotic neurons increased significantly (P < 0.001, n = 9 and 14 slices) after 12 and 24 h application of ifenprodil (Fig. 7D). However, spontaneous activity was still present in slices (n = 2) preincubated for 24 h with ifenprodil (data not shown).

Figure 6. Role of NMDA receptors and voltage-dependent calcium channels in mediating spontaneous neuronal activity in neocortical slice cultures. (A--C) MEA recordings from the same 6 div slice cultures. Eight channels are damaged and were switched off. (A) Recordings in control condition reveal spontaneous and synchronized network bursts. (B) Spontaneous activity is blocked by the NMDA receptor antagonist CPP. (C) Addition of 25 mM potassium to the CPP containing ACSF causes a transient recovery in spontaneous network activity lasting 1.5 min. The arrows indicate that artifacts are present in 2 channels. (D) Relative change in the number of apoptotic neurons compared with untreated control cultures. Blockade of the NMDA receptor with 10 lM CPP for 6 h induces a significant increase in apoptosis rate. When 10 lM CPP was combined with 25 mM extracellular potassium, the apoptosis rate was not significantly different compared with the control. Addition of 10 lM nifedipine, a specific blocker of L-type calcium channels, to the CPP and high potassium--containing solution causes an increase in the number of apoptotic cells. Nifedipine alone does not induce any change in the apoptosis rate. Kruskal--Wallis test on the apoptosis assay gave P \ 0.0001. Post hoc Dunn’s test comparing each condition against untreated control gave results as indicated. (E) The high potassium--induced rescue in CPP-treated slices is abolished in the presence of 200 nM k252a, an antagonist of the neurotrophin TrkB/C receptor. ANOVA analysis gave P \ 0.0001. Post hoc Newman--Keuls test comparing each condition gave results as indicated in the figure. (F) Measurements of phosphorylated CREB levels with ELISA assay show an increase in the presence of high potassium in comparison to control. This increase is abolished when the L-type calcium channels are blocked by 10 lM nifedipine. **P \ 0.01, ***P \ 0.001; nonsignificant (NS) 5 P [ 0.05.

Role of GABA-A Receptors in Spontaneous Network Activity and Survival The potential role of c-aminobutyric acidergic synaptic mechanisms was also assessed. In control condition, spontaneous events recorded with the MEA system showed a pattern consisting of a prominent activity in the frequency range from 5 to 8 Hz (see Fig. 3C and arrow in Fig. 8A). After treatment with 100 lM gabazine, a specific antagonist of ionotropic caminobutyric acid (GABA) receptors, the amplitude of this activity was significantly reduced (control: 142.1 ± 13.4 lV; gabazine 36.9 ± 3.1 lV; P < 0.0001 Mann--Whitney test, 15

spontaneous events from n = 3 slices, Fig. 8B). This change in activity pattern also induced a significant increase in apoptosis after 6, 12, and 24 h of 100 lM gabazine treatment (P < 0.001, n = 12 slices for 6 and 24 h; P < 0.01, n = 12 slices for 12 h). Role of Neuronal Gap-Junctions in Spontaneous Network Activity and Survival In agreement with previous observations on newborn rodent, acute slice preparations (Yuste et al. 1995; Kandler and Katz 1998) and intact cortices (Dupont et al. 2006), synchronized network activity in neocortical slice cultures from newborn Cerebral Cortex June 2008, V 18 N 6 1343

Figure 7. Role of NR2A- and NR2B-containing NMDA receptors in early neocortical network activity and activity-dependent apoptosis. (A) Blockade of NR2A-containing NMDA receptors with 400 nM NVP-AMM077 leads to a complete loss of spontaneous activity as recorded with the MEA system. (B) Application of 400 nM NVP-AMM077 for 6, 12, and 24 h induces a significant increase in the apoptosis rate (Mann--Whitney test, all P \ 0.001). Note that the apoptosis rate is higher during initial 6 h and gradually decreases between 6 and 24 h (Kruskal--Wallis test, P \ 0.0001; Dunn’s post hoc test P \ 0.001 between each condition). (C) Blockade of NR2B-containing NMDA receptors with 3 lM ifenprodil does not change spontaneous network activity. (D) Time-dependent increase in apoptosis rate in slice cultures treated with 3 lM ifenprodil. A significant increase in number of cells entering the apoptosis process can be observed after 12 and 24 h of ifenprodil treatment (Mann--Whitney test, both P \ 0.0001), but not after 6 h. Analyses with Kruskal--Wallis test confirmed the difference in the 6 versus 12- and 24-h experimental groups (Kruskal--Wallis test, P \ 0.0001; Dunn’s post hoc test P \ 0.001 between 6 h and the other periods of antagonist application).

mice required intact gap junctional coupling. Application of 25 lM mefloquine, an inhibitor of neuronal, connexin 36 containing gap junctions (Cruikshank et al. 2004), blocked the synchronized network activity (n = 5 slices, Fig. 9A,B). Interestingly, the number of action potentials recorded extracellularly with the entire MEA in a 5-min observation period increased significantly (P < 0.05) from 31.5 ± 7.2 under control conditions to 138.1 ± 42.3 in mefloquine (n = 3 slices), indicating that spontaneous activity was still present, but no longer synchronized. In agreement with these electrophysiological results, 25 lM mefloquine also caused a significant increase (P < 0.001, n = 12 for 6h and n = 18 slices for both 12 and 24 h) in the number of apoptotic neurons, but only after application for 24 h (Fig. 9C). Discussion The main conclusions of our study on the role of electrical activity in regulating caspase-3--dependent apoptosis in neonatal mouse somatosensory cortex can be summarized as follows: 1) Caspase-3--mediated apoptosis can be demonstrated in neurons of all cortical layers in vivo. In organotypic cultures from neonatal neocortex; 2) Blockade of action-potential-1344 Apoptosis in Developing Cortex

d

Heck et al.

dependent neuronal activity with TTX or blockade of ionotropic glutamate receptors induces an increase in the number of apoptotic neurons. 3) NMDA-R containing the NR2A or NR2B subunit show a different time course on the activitydependent regulation of neuronal survival. 4) Calcium influx through nifedipine-sensitive high voltage-activated calcium channels provides an additional mechanism to promote neuronal survival, which involves CREB phosphorylation and Trk receptors. 5) The inhibition of ionotropic GABA receptors induced a change in spontaneous activity that suggests a link between a specific pattern of activity and survival. 6) Gap junctional coupling promotes long-time survival. 7) Extracellular multichannel and whole-cell patch-clamp recordings reveal spontaneous synchronized network activity with a pharmacological profile that correlates with the caspase-3-dependent apoptosis. Our data demonstrate for the first time that spontaneous neuronal activity mediated by neuronal gap junctions and by glutamatergic synapses plays an important role in regulating caspase-3--dependent neuronal apoptosis in the developing cerebral cortex. Because the spontaneous electrical activity recorded in organotypic neocortical slice cultures resembles in many aspects the spontaneous activity patterns observed in vivo, our observations are probably also relevant for the

Figure 8. GABA receptor blockade causes suppression of spontaneous burst activity and increase in apoptosis. (A) Representative extracellular field potential recording with the MEA system and corresponding color-coded wavelet analysis of a 6 div neocortical slice. (B) GABA-A receptor blockade with 100 lM gabazine causes a reduction in the network activity, especially of the prominent activity in the 5- to 8Hz frequency range (arrow and white ellipse in [A] and [B]). (C) Application of 100 lM gabazine induces a significant increase in apoptosis rate after 6 h (Mann--Whitney, P \ 0.01), 12 h (Mann--Whitney, P \ 0.001), and 24 h (Mann--Whitney, P \ 0.01) of treatment.

activity-dependent regulation of neuronal apoptosis during early cortical development in vivo. Cortical Neurons Undergo Caspase-3--Dependent Apoptosis In Vivo Although several studies have assessed the occurrence of apoptosis in the developing cerebral cortex (Ferrer et al. 1992;

Figure 9. Role of neuronal gap junctions in modulating spontaneous network oscillations and activity-dependent cell death. (A) MEA recordings under control conditions reveal large-scale spontaneous and synchronized network oscillations. (B) Bath application of the neuronal gap junction blocker mefloquine blocks synchronized network activity, but leaves spontaneous nonsynchronized action potential firing preserved. (C) Blockade of neuronal gap junctions with 25 lM mefloquine induces a significant (Mann--Whitney test, P \ 0.0001) increase in apoptosis rate after 24-h treatment, but does not influence survival rate after 6- and 12-h mefloquine application (Mann--Whitney, P [ 0.05).

Cerebral Cortex June 2008, V 18 N 6 1345

Spreafico et al. 1995; Thomaidou et al. 1997; Verney et al. 2000), including human cortex (Rakic and Zecevic 2000; Chan et al. 2002), the identity of the dying cells and the type of cell death were not clear. Here, we show that caspase-3--dependent apoptosis regulates the cell number in all layers of the early postnatal cerebral cortex. An important role of caspase-3 in the control of early PCD has been previously demonstrated in mice deficient for the gene coding for caspase-3. These knockout mice show massive deficits in brain development (Kuida et al. 1996) with oversized hemispheres due to a decreased apoptosis of postmitotic neurons in the developing cerebral cortex (for review see Kuan et al. 2000). However, the early death of these mice does not allow the analysis of PCD at postnatal stages. Our in vivo analyses suggest that apoptosis does not only regulate the cell number in the embryonic proliferative zone, but also in a second wave of caspase-3--dependent PCD in the early postnatal cortex. Our results are in agreement with previous in vivo observations demonstrating a high expression of caspase-3 in the neocortex (Shimohama et al. 2001) and of the proapoptotic gene Bax (Vekrellis et al. 1997) during the early postnatal period and a subsequent decrease to low levels in the adult. In the cortex, the active form of caspase-3 has a peak of detection during the first postnatal week (Mooney and Miller 2000) and the neocortex reveals a significantly higher expression levels of Bax and active caspase-3 than other brain structures (Menshanov et al. 2006).

2005). Spontaneous network bursts could thus modify the sensitivity to death for a long period of time. This may also apply to the in vivo situation, where periods of activity are separated by long periods of silence (Khazipov and Luhmann 2006). Our data further indicate that nonsynchronized neuronal discharges are insufficient to sustain long-term survival. Gap junction antagonists blocked synchronized network activity and induced caspase-3--dependent apoptosis after 24 h although nonsynchronized action potential discharges were preserved or even enhanced under this condition. Moreover, the inhibition of GABA receptors induced a change in the activity pattern with the loss of the prominent 5- to 8-Hz component, which leads to an increase in apoptosis rate. It has been previously reported that burst activity, rather than a steady rate of single action potentials is more effective in triggering BDNF release (see e.g., Balkowiec and Katz 2000; Hartmann et al. 2001). This suggests that activity-dependent release of BDNF (for a recent review see Lessmann et al. 2003) could account for the activity-dependent neuronal survival in our cultures. These data suggest that synchronized neuronal activity or sustained depolarization is required for the activity-dependent regulation of neuronal survival in the neonatal cerebral cortex. Besides modulating spontaneous electrical activity, gap junction channels may also directly propagate cell death and survival modulating signals (for review see Krysko et al. 2005).

Neuronal Survival during Early Neocortical Development Is Activity Dependent Previous studies have demonstrated in various model systems that electrical activity plays an important role in promoting neuronal survival (for review see Mennerick and Zorumski 2000). We used organotypic slice cultures to study in more detail the mechanisms that regulate activity-dependent neuronal survival during early neocortical development. These slice cultures showed a distribution of caspase-3--dependent apoptosis that was similar to what was observed in vivo. The pattern of spontaneous neuronal activity recorded at the singlecell level with whole-cell patch-clamp recordings and at the network level with MEAs resembled the in vivo activity pattern observed in the rodent cerebral cortex during the first postnatal week (Khazipov et al. 2004; Hanganu et al. 2006) (for review see Moody and Bosma 2005; Khazipov and Luhmann 2006). Spontaneous TTX-sensitive bursts and synchronized network activity could be blocked by ionotropic glutamate receptor antagonists (Minlebaev et al. 2007) and during earlier developmental stage with gap junction blockers (Adelsberger et al. 2005; Dupont et al. 2006). When action potentials and evoked transmitter release were blocked by TTX, the number of apoptotic cells was significantly increased, demonstrating that spontaneous electrical activity is required for neuronal survival. Our results are in good agreement with previous observations of massive cell loss in knockout mice deficient for voltage-gated sodium channels (Planells-Cases et al. 2000) and for munc13 (Verhage et al. 2000). The pattern of activity that leads to survival was characterized by spontaneous network events occurring at long time intervals, indicating that a permanent activity level is not required for survival. It has been recently demonstrated that an episode of high activity prior to exposure to an apoptotic stimulus has a long-term neuroprotective effect (Papadia et al.

Role of Glutamatergic Receptors and Postynaptic Calcium Level The action of this anti-apoptotic activity code requires activation of ionotropic glutamate receptors. Blockade of AMPA/kainate as well as NMDA-R caused a 2- to 2.5-fold increase in the apoptosis rate. An essential role of NMDA-R in activity-dependent survival during early cortical development has been previously demonstrated in culture (Hwang et al. 1999; Yoon et al. 2003) and in vivo (Ikonomidou et al. 1999; Tome et al. 2006). NMDA-R deficient mice also reveal an approximately 2-fold increase in developmental cell death, which depends on caspase-3 and Bax (Adams et al. 2004; Rivero Vaccari et al. 2006). Our data demonstrate for the first time a contribution of different NMDA-R subunits in the control of caspase-3 dependent apoptosis. Interestingly, NR2A and NR2B subunits seem to play a different role in activity-dependent survival. Blockade of NR2A-containing NMDA-Rs causes an immediate increase in PCD, which gradually declines over 12-24 h. In contrast, blockade of the NR2B subunit induces a delayed onset in apoptosis which remains constant over at least 24 h. Activation of NMDA-Rs promotes a short-term as well as a long-lasting neuroprotection, both mediated by independent pathways (Papadia et al. 2005; Soriano et al. 2006). The NR2B subunit is specifically coupled to the ERK pathway (Krapivinsky et al. 2003) and extracellular signalregulated kinase (ERK) is involved in long-lasting neuroprotection by regulating CREB phosphorylation (Hardingham and Bading 2003; Papadia et al. 2005). Therefore, we postulate on the basis of our results that long-lasting activity-dependent cell survival is mediated by NR2B, whereas short-term survival depends on NR2A. Further support for this hypothesis comes from recent studies (Zhou and Baudry 2006; Chen et al. 2007), which demonstrate distinct roles for NR2A and NR2B in neuroprotection.

1346 Apoptosis in Developing Cortex

d

Heck et al.

Our data indicate that high-threshold L-type calcium channels are also capable to activate a prosurvival pathway. Under blockade of NMDA-R, an increase in [K+]e elicited a transient synchronized network activity that was sufficient to prevent the CPP-induced apoptosis. Nifedipine blocked this high [K+]e-induced prosurvival process but did not influence the apoptosis rate under normal experimental conditions. These data clearly indicate that an intracellular calcium rise via L-type calcium channels represents an alternative survival pathway (Ghosh et al. 1994; Tao et al. 1998). Although calcium influx via NMDA-Rs and L-type calcium channels activate different calcium-binding proteins (Marshall et al. 2003), both pathways activate kinases leading to CREB dependent BDNF synthesis (Ghosh et al. 1994; Dolmetsch et al. 2001). In our model, the activation of L-type calcium channels was associated with an increase in the level of P-CREB and the rescue of the neurons from the blockade of NMDA-R was dependent on TrkB/C receptors, which are activated by BDNF and NT-3. Our results demonstrate that in a physiological neuronal network both pathways can complement each other. This supports the hypothesis that postsynaptic calcium currents are one essential component of the survival (for review see Hara and Snyder 2007). Interestingly, Ca2+-influx via NMDA-Rs or L-type voltagegated calcium channels are the 2 key mechanisms to elicit activity-dependent release of BDNF and NT-3 in central neuronal cultures (Kolarow et al. 2007). It is thus tempting to speculate that release of neurotrophins driven by network activity is a possible link between activity and neuronal survival in our model. Conclusion We demonstrate that during early development, neocortical neurons undergo activity-dependent, caspase-3--mediated apoptosis. Synchronized spontaneous network bursts activating ionotropic glutamate receptors and high-threshold calcium channels as well as gap junctional coupling promotes neuronal survival. NMDA-Rs containing NR2A and NR2B subunits show a differential and time-dependent influence on the regulation of apoptosis indicating that different calcium-dependent pathways are activated. Our data suggest that disturbances of early network activity also interfere with the activity-dependent regulation of developmental apoptosis causing long-term structural and functional alterations in neocortical circuitry. Funding Schram-Stiftung (T287/16255/2006 to V. L. and H. J. L.). Notes We thank Beate Krumm for excellent technical assistance, Dr Werner Kilb for help with the electrophysiological experiments, and Dr Tanja Brigadski and Richard Kolarow for help on the organotypic slice cultures. AG and JJS are members of the neuroscience graduate school at the University of Mainz (DFG GRK 1044). Conflict of Interest: None declared. Address correspondence to email: [email protected].

References Adams SM, Rivero Vaccari JC, Corriveau RA. 2004. Pronounced cell death in the absence of NMDA receptors in the developing somatosensory thalamus. J Neurosci. 24:9441--9450. Adelsberger H, Garaschuk O, Konnerth A. 2005. Cortical calcium waves in resting newborn mice. Nat Neurosci. 8:988--990.

Balkowiec A, Katz DM. 2000. Activity-dependent release of endogenous brain-derived neurotrophic factor from primary sensory neurons detected by ELISA in situ. J Neurosci. 20:7417--7423. Banks GB, Kanjhan R, Wiese S, Kneussel M, Wong LM, O’Sullivan G, Sendtner M, Bellingham MC, Betz H, Noakes PG. 2005. Glycinergic and GABAergic synaptic activity differentially regulate motoneuron survival and skeletal muscle innervation. J Neurosci. 25:1249--1259. Benn SC, Woolf CJ. 2004. Adult neuron survival strategy—slamming on the brakes. Nat Rev Neurosci. 5:686--700. Berberich S, Punnakkal P, Jensen V, Pawlak V, Seeburg PH, Hvalby O, Kohr G. 2005. Lack of NMDA receptor subtype selectivity for hippocampal long-term potentiation. J Neurosci. 25:6907--6910. Brewer GJ, Torricelli JR, Evege EK, Price PJ. 1993. Optimized survival of hippocampal neurons in B27-supplemented neurobasal, a new serum-free medium combination. J Neurosci Res. 35:567--576. Buss RR, Sun W, Oppenheim RW. 2006. Adaptive roles of programmed cell death during nervous system development. Annu Rev Neurosci. 29:1--35. Chan WY, Lorke DE, Tiu SC, Yew DT. 2002. Proliferation and apoptosis in the developing human neocortex. Anat Rec. 267:261--276. Chen Q, He S, Hu XL, Yu J, Zhou Y, Zheng J, Zhang S, Zhang C, Duan WH, Xiong ZQ. 2007. Differential roles of NR2A- and NR2B-containing NMDA receptors in activity-dependent brain-derived neurotrophic factor gene regulation and limbic epileptogenesis. J Neurosci. 27:542--552. Corlew R, Bosma MM, Moody WJ. 2004. Spontaneous, synchronous electrical activity in neonatal mouse cortical neurons. J Physiol. 560:377--390. Cruikshank SJ, Hopperstad M, Younger M, Connors BW, Spray DC, Srinivas M. 2004. Potent block of Cx36 and Cx50 gap junction channels by mefloquine. Proc Natl Acad Sci USA. 101:12364--12369. Davies AM. 2003. Regulation of neuronal survival and death by extracellular signals during development. EMBO J. 22:2537--2545. Dolmetsch RE, Pajvani U, Fife K, Spotts JM, Greenberg ME. 2001. Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway. Science. 294:333--339. Dupont E, Hanganu IL, Kilb W, Hirsch S, Luhmann HJ. 2006. Rapid developmental switch in the mechanisms driving early cortical columnar networks. Nature. 439:79--83. Ferrer I, Soriano E, Del Rio JA, Alcantara S, Auladell C. 1992. Cell death and removal in the cerebral cortex during development. Prog Neurobiol. 39:1--43. Galli-Resta L, Resta G. 1992. A quantitative model for the regulation of naturally occurring cell death in the developing vertebrate nervous system. J Neurosci. 12:4586--4594. Ghosh A, Carnahan J, Greenberg ME. 1994. Requirement for BDNF in activity-dependent survival of cortical neurons. Science. 263: 1618--1623. Hanganu IL, Ben Ari Y, Khazipov R. 2006. Retinal waves trigger spindle bursts in the neonatal rat visual cortex. J Neurosci. 26:6728--6736. Hansen HH, Briem T, Dzietko M, Sifringer M, Voss A, Rzeski W, Zdzisinska B, Thor F, Heumann R, Stepulak A, et al. 2004. Mechanisms leading to disseminated apoptosis following NMDA receptor blockade in the developing rat brain. Neurobiol Dis. 16:440--453. Hara MR, Snyder SH. 2007. Cell signaling and neuronal death. Annu Rev Pharmacol Toxicol. 47:117--141. Hardingham GE, Bading H. 2003. The Yin and Yang of NMDA receptor signalling. Trends Neurosci. 26:81--89. Hardingham GE, Fukunaga Y, Bading H. 2002. Extrasynaptic NMDARs oppose synaptic NMDARs by triggering CREB shut-off and cell death pathways. Nat Neurosci. 5:405--414. Hartmann M, Heumann R, Lessmann V. 2001. Synaptic secretion of BDNF after high-frequency stimulation of glutamatergic synapses. EMBO J. 20:5887--5897. Haydar TF, Bambrick LL, Krueger BK, Rakic P. 1999. Organotypic slice cultures for analysis of proliferation, cell death, and migration in the embryonic neocortex. Brain Res Protoc. 4:425--437. Heumann D, Leuba G. 1983. Neuronal death in the development and aging of the cerebral cortex of the mouse. Neuropathol Appl Neurobiol. 9:297--311. Heumann D, Leuba G, Rabinowicz T. 1978. Postnatal development of the mouse cerebral neocortex. IV. Evolution of the total cortical

Cerebral Cortex June 2008, V 18 N 6 1347

volume, of the population of neurons and glial cells. J Hirnforsch. 19:385--393. Hidalgo A, Ffrench-Constant C. 2003. The control of cell number during central nervous system development in flies and mice. Mech Dev. 120:1311--1325. Hwang JY, Kim YH, Ahn YH, Wie MB, Koh JY. 1999. N-Methyl-Daspartate receptor blockade induces neuronal apoptosis in cortical culture. Exp Neurol. 159:124--130. Ikonomidou C, Bosch F, Miksa M, Bittigau P, Vockler J, Dikranian K, Tenkova TI, Stefovska V, Turski L, Olney JW. 1999. Blockade of NMDA receptors and apoptotic neurodegeneration in the developing brain. Science. 283:70--74. Innocenti GM, Price DJ. 2005. Exuberance in the development of cortical networks. Nat Rev Neurosci. 6:955--965. Kandler K, Katz LC. 1998. Coordination of neuronal activity in developing visual cortex by gap junction-mediated biochemical communication. J Neurosci. 18:1419--1427. Khazipov R, Luhmann HJ. 2006. Early patterns of electrical activity in the developing cerebral cortex of humans and rodents. Trends Neurosci. 29:414--418. Khazipov R, Sirota A, Leinekugel X, Holmes GL, Ben Ari Y, Buzsaki G. 2004. Early motor activity drives spindle bursts in the developing somatosensory cortex. Nature. 432:758--761. Kilb W, Luhmann HJ. 2003. Carbachol induced network oscillations in the intact cerebral cortex of the newborn rat. Cereb Cortex. 13:409--421. Kolarow R, Brigadski T, Lessmann V. 2007. Postsynaptic secretion of BDNF and NT-3 from hippocampal neurons depends on calciumcalmodulin kinase II signaling and proceeds via delayed fusion pore opening. J Neurosci. 27(39):10350--10364. Krapivinsky G, Krapivinsky L, Manasian Y, Ivanov A, Tyzio R, Pellegrino C, Ben Ari Y, Clapham DE, Medina I. 2003. The NMDA receptor is coupled to the ERK pathway by a direct interaction between NR2B and RasGRF1. Neuron. 40:775--784. Krysko DV, Leybaert L, Vandenabeele P, D’Herde K. 2005. Gap junctions and the propagation of cell survival and cell death signals. Apoptosis. 10:459--469. Kuan CY, Roth KA, Flavell RA, Rakic P. 2000. Mechanisms of programmed cell death in the developing brain. Trends Neurosci. 23:291--297. Kuida K, Zheng TS, Na S, Kuan C, Yang D, Karasuyama H, Rakic P, Flavell RA. 1996. Decreased apoptosis in the brain and premature lethality in CPP32-deficient mice. Nature. 384:368--372. Lavoie JF, Lesauteur L, Kohn J, Wong J, Furtoss O, Thiele CJ, Miller FD, Kaplan DR. 2005. TrkA induces apoptosis of neuroblastoma cells and does so via a p53-dependent mechanism. J Biol Chem. 280:29199--29207. Lessmann V, Gottmann K, Malcangio M. 2003. Neurotrophin secretion: current facts and future prospects. Prog Neurobiol. 69: 341--374. Linden R. 1994. The survival of developing neurons: a review of afferent control. Neuroscience. 58:671--682. Liot G, Gabriel C, Cacquevel M, Ali C, MacKenzie ET, Buisson A, Vivien D. 2004. Neurotrophin-3-induced PI-3 kinase/Akt signaling rescues cortical neurons from apoptosis. Exp Neurol. 187:38--46. Luhmann HJ, Reiprich RA, Hanganu IL, Kilb W. 2000. Cellular physiology of the neonatal rat cerebral cortex: intrinsic membrane properties, sodium and calcium currents. J Neurosci Res. 62:574--584. Marin-Teva JL, Dusart I, Colin C, Gervais A, van Rooijen N, Mallat M. 2004. Microglia promote the death of developing Purkinje cells. Neuron. 41:535--547. Marshall J, Dolan BM, Garcia EP, Sathe S, Tang X, Mao Z, Blair LA. 2003. Calcium channel and NMDA receptor activities differentially regulate nuclear C/EBPbeta levels to control neuronal survival. Neuron. 39:625--639. Mennerick S, Zorumski CF. 2000. Neural activity and survival in the developing nervous system. Mol Neurobiol. 22:41--54. Menshanov PN, Bannova AV, Dygalo NN. 2006. Region-specific interrelations between apoptotic proteins expression and DNA

1348 Apoptosis in Developing Cortex

d

Heck et al.

fragmentation in the neonatal rat brain. Neurochem Res. 31:869--875. Minlebaev M, Ben Ari Y, Khazipov R. 2007. Network mechanisms of spindle-burst oscillations in the neonatal rat barrel cortex in vivo. J Neurophysiol. 97:692--700. Moody WJ, Bosma MM. 2005. Ion channel development, spontaneous activity, and activity-dependent development in nerve and muscle cells. Physiol Rev. 85:883--941. Mooney SM, Miller MW. 2000. Expression of bcl-2, bax, and caspase-3 in the brain of the developing rat. Brain Res Dev Brain Res. 123: 103--117. Oppenheim RW. 1991. Cell death during development of the nervous system. Annu Rev Neurosci. 14:453--501. Papadia S, Stevenson P, Hardingham NR, Bading H, Hardingham GE. 2005. Nuclear Ca2+ and the cAMP response element-binding protein family mediate a late phase of activity-dependent neuroprotection. J Neurosci. 25:4279--4287. Planells-Cases R, Caprini M, Zhang J, Rockenstein EM, Rivera RR, Murre C, Masliah E, Montal M. 2000. Neuronal death and perinatal lethality in voltage-gated sodium channel alpha(II)-deficient mice. Biophys J. 78:2878--2891. Posada A, Clarke PG. 1999. The role of neuronal death during the development of topographically ordered projections: a computational approach. Biol Cybern. 81:239--247. Poser S, Impey S, Xia Z, Storm DR. 2003. Brain-derived neurotrophic factor protection of cortical neurons from serum withdrawalinduced apoptosis is inhibited by cAMP. J Neurosci. 23:4420--4427. Rakic S, Zecevic N. 2000. Programmed cell death in the developing human telencephalon. Eur J Neurosci. 12:2721--2734. Rivero Vaccari JC, Casey GP, Aleem S, Park WM, Corriveau RA. 2006. NMDA receptors promote survival in somatosensory relay nuclei by inhibiting Bax-dependent developmental cell death. Proc Natl Acad Sci USA. 103:16971--16976. Sendtner M, Pei G, Beck M, Schweizer U, Wiese S. 2000. Developmental motoneuron cell death and neurotrophic factors. Cell Tissue Res. 301:71--84. Shimohama S, Tanino H, Fujimoto S. 2001. Differential expression of rat brain caspase family proteins during development and aging. Biochem Biophys Res Commun. 289:1063--1066. Soriano FX, Papadia S, Hofmann F, Hardingham NR, Bading H, Hardingham GE. 2006. Preconditioning doses of NMDA promote neuroprotection by enhancing neuronal excitability. J Neurosci. 26:4509--4518. Spreafico R, Frassoni C, Arcelli P, Selvaggio M, De Biasi S. 1995. In situ labeling of apoptotic cell death in the cerebral cortex and thalamus of rats during development. J Comp Neurol. 363:281--295. Stoppini L, Buchs PA, Muller D. 1991. A simple method for organotypic cultures of nervous tissue. J Neurosci Methods. 57:985--994. Tao X, Finkbeiner S, Arnold DB, Shaywitz AJ, Greenberg ME. 1998. Ca2+ influx regulates BDNF transcription by a CREB family transcription factor-dependent mechanism. Neuron. 20:709--726. Thomaidou D, Mione MC, Cavanagh JF, Parnavelas JG. 1997. Apoptosis and its relation to the cell cycle in the developing cerebral cortex. J Neurosci. 17:1075--1085. Tome CM, Bauer C, Nottingham C, Smith C, Blackstone K, Brown L, Hlavaty C, Nelson C, Daker R, Sola R, et al. 2006. Mk801-induced caspase-3 in the postnatal brain: inverse relationship with calcium binding proteins. Neuroscience. 141:1351--1363. Vekrellis K, McCarthy MJ, Watson A, Whitfield J, Rubin LL, Ham J. 1997. Bax promotes neuronal cell death and is downregulated during the development of the nervous system. Development. 124: 1239--1249. Verhage M, Maia AS, Plomp JJ, Brussaard AB, Heeroma JH, Vermeer H, Toonen RF, Hammer RE, van den Berg TK, Missler M, et al. 2000. Synaptic assembly of the brain in the absence of neurotransmitter secretion. Science. 287:864--869. Verney C, Takahashi T, Bhide PG, Nowakowski RS, Caviness VS Jr. 2000. Independent controls for neocortical neuron production and histogenetic cell death. Dev Neurosci. 22:125--138. Voyvodic JT. 1996. Cell death in cortical development: how much? Why? So what? Neuron. 16:693--696.

Walz C, Jungling K, Lessmann V, Gottmann K. 2006. Presynaptic plasticity in an immature neocortical network requires NMDA receptor activation and BDNF release. J Neurophysiol. 96:3512--3516. Williams K. 1993. Ifenprodil discriminates subtypes of the Nmethyl-D-aspartate receptor: selectivity and mechanisms at recombinant heteromeric receptors. Mol Pharmacol. 44:851--859. Yoon WJ, Won SJ, Ryu BR, Gwag BJ. 2003. Blockade of ionotropic glutamate receptors produces neuronal apoptosis through the Bax-

cytochrome C-caspase pathway: the causative role of Ca2+ deficiency. J Neurochem. 85:525--533. Yuste R, Nelson DA, Rubin WW, Katz LC. 1995. Neuronal domains in developing neocortex: mechanisms of coactivation. Neuron. 14:7--17. Zhou M, Baudry M. 2006. Developmental changes in NMDA neurotoxicity reflect developmental changes in subunit composition of NMDA receptors. J Neurosci. 26:2956--2963.

Cerebral Cortex June 2008, V 18 N 6 1349