Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Aug 1.
Published in final edited form as: J Neurochem. 2010 May 28;114(4):1107–1118. doi: 10.1111/j.1471-4159.2010.06835.x

NR2B-NMDA receptor mediated increases in intracellular Ca2+ concentration regulate the tyrosine phosphatase, STEP, and ERK MAP kinase signaling

Surojit Paul *,§,, John A Connor §
PMCID: PMC3049732  NIHMSID: NIHMS267918  PMID: 20524968

Abstract

NMDA receptors regulate both the activation and inactivation of the extracellular signal-regulated kinase (ERK) signaling cascade, a key pathway involved in neuronal plasticity and survival. This bi-directional regulation of ERK activity by NMDA receptors has been attributed to opposing actions of NR2A- versus NR2B-containing NMDA receptors but how this is implemented is not understood. Here we show that glutamate-mediated intracellular Ca2+ increases occur in two phases, a rapid initial increase followed by a delayed larger increase. Both phases of the Ca2+ increase were blocked by MK-801, a non-selective NMDA receptor inhibitor. On the other hand selective inhibition of NR2B-NMDA receptors by Ifenprodil or Ro 25-6981 blocked the delayed larger phase but had only a small effect on the rapid initial increase. The rapid initial increase in Ca2+, presumably due to NR2A-NMDAR activation, was sufficient to activate ERK, whereas the large delayed increases in Ca2+ mediated by NR2B-NMDARs were necessary for dephosphorylation and subsequent activation of STEP, a neuron-specific tyrosine phosphatase that in turn mediates the dephosphorylation and inactivation of ERK. We conclude that the magnitude of Ca2+ increases mediated through NR2B-NMDA receptors plays a critical role in the regulation of the serine/threonine and tyrosine kinases and phosphatases that are involved in the regulation of ERK activity.

Keywords: ERK, STEP, NMDA, calcium, tyrosine phosphatase

INTRODUCTION

The tyrosine phosphatase STEP (striatal-enriched phosphatase) is a signaling molecule downstream of both dopamine and NMDA receptors (Paul et al. 2000; Paul et al. 2003). STEP is highly enriched within the medium spiny neurons of the basal ganglia and related structures (Lombroso et al. 1993; Boulanger et al. 1995). There are multiple splice variants of STEP that include both cytosolic (STEP46) and membrane-associated (STEP61) variants (Sharma et al. 1995; Bult et al. 1997). The ability of STEP to bind to its substrates is regulated by phosphorylation of a serine residue located within the kinase-interacting motif or KIM domain. Phosphorylation of this serine residue in response to activation of dopamine/D1 receptors/cAMP-dependent protein kinase-A pathway, decreases the activity of STEP by reducing its affinity for its substrates (Paul et al. 2000). Dephosphorylation of this same serine residue activates STEP and has been shown to occur via calcineurin activity promoted by glutamate/NMDA receptor-mediated Ca2+ influx (Paul et al. 2003). Both in vitro and in vivo studies show that the activity of STEP is low in neurons under basal conditions (Paul et al. 2003; Valjent et al. 2005). Once activated, STEP can bind to and down-regulate the activity of extracellular-regulated kinase 2 (ERK2), a key signaling protein involved in NMDA receptor dependent synaptic plasticity and excitotoxicity (Komiyama et al. 2002; Banko et al. 2004; Thomas and Huganir 2004; Kaphzan et al. 2006; Paul et al. 2007). NMDA receptor stimulation by glutamate or NMDA leads to rapid but transient phosphorylation of ERK2. Both phosphorylation and dephosphorylation of ERK2 is Ca2+ dependent (Jiang et al. 2000b, a; Chandler et al. 2001; Paul et al. 2003). The dephosphorylation of ERK2 has been attributed to the delayed activation of STEP following NMDA receptor stimulation (Paul et al. 2003) but the mechanisms by which the same messenger, Ca2+, is signaling both the phosphorylation and dephosphorylation of ERK2 has not been revealed.

The aim of the present study was to determine the relative contribution of NR2A- and NR2B-NMDARs in mediating intracellular Ca2+ increases and examine if the extent of Ca2+ increase is a critical determinant of STEP and ERK activity in neurons. Our results show that NMDA receptor stimulation leads to an initial rapid rise in Ca2+ followed by a gradual but larger increase in Ca2+. The delayed increase in Ca2+ is NR2B-NMDAR-dependent. We also show that glutamate-mediated dephosphorylation and subsequent activation of STEP is dependent on this large delayed, NR2B-NMDAR-dependent increase. The preferential activation of STEP by NR2B explains the sub-type specific function of NR2B-NMDARs in the inhibition of ERK activity.

MATERIALS AND METHODS

Materials and reagents

Pregnant female Sprague-Dawley rats (16 day gestation) were from Harlan Laboratories. Antibodies used were as follows: polyclonal anti-NR1 from Abcam, polyclonal anti-NR2B from Millipore, polyclonal anti-NR2A from Covance Research Products, monoclonal anti-STEP from Novus Biologicals, polyclonal anti-phospho-ERK1/2 (TEYP) from Santa Cruz Biotechnology and polyclonal anti-tubulin from Sigma-Aldrich. All secondary antibodies, Fura 2-AM and Fura FF-AM were from Invitrogen. All other reagents were from Sigma-Aldrich. Approval for animal experiments was given by the University of New Mexico, Health Sciences center, Institutional Animal Care and Use Committee.

Cell culture and stimulation

Primary neuronal cultures were obtained from 16–17 day old rat embryos as described previously (Paul et al. 2003). Briefly, the striatum and the adjoining cortex were dissected, the tissue dissociated mechanically and resuspended in DMEM/F-12 (1:1)-containing 5% fetal calf serum. Cells were plated on poly-D-lysine-coated tissue culture dishes (6 × 106 cells/dish) and grown for 12–14 days at 37°C in a humidified atmosphere (95:5% air:CO2 mixture). To inhibit proliferation of non-neuronal cells, 10 μM of cytosine D-arabinofuranoside was added to the cultures 72 hr after plating. For receptor stimulation cells were treated with glutamate, NMDA or KCl for the indicated times at 37°C. 1 μM of glycine was added to the media during treatment with glutamate or NMDA. For biochemical experiments NMDA receptor antagonists MK-801, Ifenprodil or Ro 25-6981 (Fischer et al. 1997; Williams 2001; Waxman and Lynch 2005; Kohr 2006) was added 20 min prior to stimulation. The doses selected for MK-801, Ifenprodil and Ro 25-6981 were based on previous studies that have used the compounds to determine the role of NR2B-NMDA receptors in the regulation of NMDA receptor currents, protein phosphorylation, downstream signaling cascades as well as neuronal cell death (Crozier et al. 1999; Tovar and Westbrook 1999; Kim et al. 2005; Waxman and Lynch 2005; Yang et al. 2006; Martel et al. 2009). For Ca2+ measurement experiments these antagonist were added at the times indicated.

Immunoprecipitation

Cells were lysed in a buffer containing 50 mM Tris-HCl, pH 7.4, 150 mM NaCl, 50 mM NaF, 10 mM Na4P2O7, 1 mM Na3VO4, 0.1% NP-40, 0.1% sodium-deoxycholate and protease inhibitor cocktail (Boehringer). Lysates were centrifuged for 10 min at 14,000 rpm to remove insoluble material, and then pre-cleared with protein G-sepharose for 1 hr. For immunoprecipitation of STEP, samples were incubated overnight with anti-STEP antibody. The immune complexes were incubated with 30 μl of protein G-sepharose for 2 hr at 4°C. Beads were collected by centrifugation at 1000 rpm for 2 min and washed five times with the lysis buffer. Proteins were eluted using SDS-sample buffer and processed for immunoblot analysis with antibodies as described in the individual experiments. Densitometric analysis of immunoblots was performed using Image J software (NIH, Bethesda, MD, USA). Statistical comparison was carried out using One-way analysis of variance (ANOVA, Bonfrroni's multiple comparison test) and differences were considered significant when p < 0.05.

Immunofluorescence

To demonstrate co-expression of STEP and NMDA receptor subunits in neurons, 12–14 days old neuronal cultures grown on poly-D-lysine coated 2-well chamber slides (BD Biosciences) were fixed with 4% para-formaldehyde in phosphate-buffered saline (PBS, pH 7.4) for 10 min, permeabilized with 0.1% Triton-X-100 in PBS for 10 min, blocked with 10% normal goat serum and 1% BSA in PBS for 1 hr followed by incubation with anti-STEP antibody and anti-NR1, -NR2A or -NR2B antibody overnight at 4°C. Cells were then incubated with anti-mouse Alexa 488 and anti-rabbit Cy3 conjugated secondary antibodies, followed by washing with PBS and mounting on glass slides. Imaging was performed with a Zeiss Axiovert 200M fluorescence microscope with attached AxioCam CCD camera using 20× and 63× objective lenses.

Calcium Measurements

Neuronal cultures were grown on poly-D-lysine coated 2-well chambered glass slides (Nunc, New York) for 12 – 14 days. Cytosolic Ca2+ determination was performed using 1 of the 2 fluorescent indicators, Fura-2/AM (Kd ~ 0.22 μM) or the lower affinity indicator Fura-FF/AM (Kd ~ 5.5 μM, both from Molecular Probes). Cells were placed in MEM (without glutamic acid and phenol red), containing either Fura-2/AM (5 μM) or Fura-FF/AM (5 μM) for a 30 min loading time. Bathing medium was then changed to one without indicator and cells were given a 20 min post incubation time at 37°C (Connor 1986). Covered culture plates were then placed on temperature-controlled stage (33°C) of a Zeiss IM35 microscope and continuously gassed with pre-moistened 95:5% Air:CO2 mixture. Fields of 5–20 cells were viewed from underneath using a 40× glycerin objective (Nikon). Excitation light was obtained from a 100 W HBO short arc lamp through 340 and 380 nm narrowband interference filters mounted in a Sutter Instruments Wheel. The standard exposure time was 200 ms for each wavelength. Fluorescence emissions (> 500 nm) were captured using a Photometrices Pixel, cooled CCD, camera (512 × 512 elements). Maximum frame pair rate in was ~ 2/sec. The microscope was also fitted with a Ludl stepper motor stage that allowed precise repositioning of recording locations for multisite recordings during responses. IP Lab software running on a Macintosh G3 computer was used for data acquisition and analysis. Ratio data from regions of interest (neuron somata and processes) were extracted from background-subtracted image pairs (340 and 380 nm excitation). Ratios were converted into estimates of free intracellular Ca2+ concentration as described previously (Connor 1986) using the equation developed by Grynkiewicz et al., (Grynkiewicz et al. 1985). Minimum and maximum fluorescence ratios determined by bath indicator loaded cells in saline containing 2 μM ionomycin and either 0 Ca2+, 0.5 mM EGTA or 2 mM Ca2+.

RESULTS

NR2B-NMDA receptor mediated dephosphorylation of STEP and ERK2

Initial studies examined the expression profile of STEP and NMDA receptor subunits in cultured neurons (12–14 days in vitro) by immunocytochemical staining with antibodies specific for STEP, NR1, NR2A or NR2B (Fig. 1A). As previously reported, STEP is expressed in both the soma and the processes of all neurons in the cultures (Lombroso et al. 1993; Paul et al. 2003). NR1 subunits are heavily expressed in the neuronal processes but are also present in the soma. NR2A and NR2B subunits are abundant in both the cell bodies and the neuronal processes, which is in agreement with previous studies (Petralia et al. 1994; Paquet et al. 1997; Rao et al. 1998; Charton et al. 1999; Tongiorgi et al. 1999; Hallett et al. 2006). Co-expression of STEP and NMDA receptor subunits in the same cells is demonstrated by overlaying the images produced with anti-STEP and anti-NR1, -NR2A or -NR2B antibodies (Fig. 1A, right column). It is also evident from the merged images of the high magnification photomicrographs of neuronal processes that both STEP and NMDA receptor subunits are co-expressed in the same processes. To determine if STEP interacts with the NMDA receptor complex in cultured neurons we immunoprecipitated STEP from neuronal lysates using anti-STEP antibody. Since all functional NMDA receptors are composed of NR1 subunits, co-immunoprecipitation of NMDA receptor complexes with STEP was analyzed using anti-NR1 antibody. The results show that STEP61, the membrane bound isoform of the STEP-family led to co-precipitation of NR1 (Fig. 1B). This is consistent with earlier findings showing that STEP61 is a component of the NMDA receptor complex in vivo (Pelkey et al. 2002; Braithwaite et al. 2006).

Figure 1.

Figure 1

Co-expression of STEP and NMDAR subunits in neurons and NR2B-NMDA receptor-mediated dephosphorylation of STEP and ERK2. (A) Immunocytochemical analysis showing the co-expression of STEP and NMDA receptor subunits NR1, NR2A and NR2B in both neuronal somata and processes. High magnification (63×) images of neuronal processes are shown in the horizontal strips below each lower resolution (20×) panel (B) Immunoblot analysis demonstrating that STEP is part of the NMDAR complex. STEP61 was immunoprecipitated from neuronal lysates with anti-STEP antibody, analyzed by SDS-PAGE and probed with anti-NR1 antibody (upper panel). The blot was re-probed with STEP antibody (lower panel). Immunoprecipitation with Protein G-Sepharose beads alone was included as a control. To ensure the presence of both STEP and NR1 in the total lysate (input lysate used for immunoprecipitation), equal amount of protein from the lysates was processed for immunoblot analysis with anti-NR1 and -STEP antibodies. (C) Immunoblot analysis demonstrating the dephosphorylation of STEP following NR2B-NMDAR activation. Neurons were treated with glutamate for 30 min, in the presence or absence of Ro 25–6981 or Ifenprodil. Equal amount of protein from each sample were analyzed by SDS-PAGE and probed with anti-STEP antibody (upper panel). Total ERK2 was also analyzed to indicate total protein loading (lower panel). (D) STEP is endogenously phosphorylated and glutamate exposure leads to its dephosphorylation. Neurons were treated with or without glutamate for 30 min and STEP61 was immunoprecipitated with anti-STEP antibody. The samples were then processed for immunoblotting with the phospho-specific STEP antibody (left panel). The blot was then re-probed with anti-STEP antibody (right panel). (E) Dephosphorylation of ERK2 correlates with activation of STEP. Neuronal cultures were treated with glutamate for the specified time periods in the presence or absence of Ro 25–6981. Phosphorylated ERK2 was identified using a phospho-specific antibody that recognizes ERK1/2 only when phosphorylated at the regulatory tyrosine residue (TEYPERK, upper panel). Total ERK2 and STEP was analyzed by probing the membrane with either anti-ERK2 antibody (middle panel) or anti-STEP antibody (lower panel). Quantification of phosphorylated ERK2 (active ERK) and dephosphorylated STEP (active STEP) was done by computer-assisted densitometry. Values are mean ± SEM (n = 3). *Indicates significant difference from 5 min glutamate treatment (p < 0.001). #Indicates significant difference from 0 min time point (p < 0.001).

Previous studies showed that glutamate-mediated NMDA receptor stimulation leads to dephosphorylation and subsequent activation of STEP (Paul et al. 2003). To examine the respective role of NR2A- and NR2B-containing NMDA receptors in the dephosphorylation of STEP we treated neurons with glutamate (100 μM, glycine 1 μM) for 30 min in the absence or presence of NR2B-NMDAR antagonists ifenprodil and Ro 25-6981. Cell lysates were then analyzed by immunoblotting with anti-STEP antibody. Under basal condition STEP61 was present as a doublet and the upper band (phosphorylated form) constituted the principal component of the doublet (Fig. 1C, lanes 1 and 5). Incubation of neurons with glutamate alone resulted in a marked decrease in the upper band and the appearance of a predominant lower band, indicating that STEP61 is rapidly dephosphorylated following glutamate treatment (Fig. 1C, lanes 2 and 6). Ro 25-6981 and Ifenprodil both blocked the downward shift in the mobility of the upper band, supporting a role for NR2B-NMDAR functioning in the dephosphorylation of STEP61 (Fig. 1C, lanes 3 and 7). Using a phospho-specific antibody that recognizes STEP only when it is phosphorylated at the protein kinase A site in the KIM domain (Paul et al. 2003) we further demonstrated that STEP61 is basally phosphorylated at this site. A marked decrease in this basal phosphorylation of STEP61 was evident following treatment with glutamate (Fig. 1D, left panel). Re-probing the membrane with anti-STEP antibody showed that glutamate treatment resulted in significant loss of the upper band that was detected with phospho-specific antibody and the appearance of the lower dephosphorylated form of STEP61 (Fig. 1D, right panel).

To examine the role of NR2B-NMDARs in the regulation of ERK2 phosphorylation, neuronal cultures were treated with glutamate (100 μM) in the presence or absence of Ro 25-6981. Immunoblot analysis with the phospho-specific ERK antibody, that recognizes only the tyrosine phosphorylated ERK1/2 (TEYP-ERK), showed that glutamate stimulated the tyrosine phosphorylation of ERK2 within 5 min with a reduction to near basal levels occurring by 30 min (Fig. 1E, upper panel, lanes 2 and 4). Incubation with Ro 25-6981 prior to treatment with glutamate had no significant effect on the phosphorylation of ERK2 at 5 min (Fig. 1E, upper panel, lane 3) but prevented the decrease in phosphorylation of ERK2 that is normally observed by 30 min (Fig. 1E, upper panel, lane 5). Analyzing the same samples with anti-STEP antibody showed that incubation with Ro 25-6981 also blocked the glutamate-mediated dephosphorylation of STEP (Fig. 1E, lower panel, lane 5) by 30 min. Thus glutamate-mediated NR2B-NMDARs stimulation is responsible for the dephosphorylation and activation of STEP and subsequent tyrosine dephosphorylation of ERK2. After exposure to Ro 25-6981 there was a trend towards decreased active STEP at 5 min from basal levels, which correlates with the trend to increased active ERK at the same time point.

NR2B-NMDAR dependent changes in intracellular Ca2+ concentration

Since NR2B-NMDAR antagonists blocked the dephosphorylation of STEP it was important to determine whether these antagonists also produced a significant change in intracellular Ca2+ levels in response to glutamate. As glutamate-induced Ca2+ increases are often very large (Petrozzino et al. 1995; Hyrc et al. 1997; Cheng et al. 1999; Stout and Reynolds 1999), we used the low affinity indicator Fura-FF for most of these experiments. Figure 2A shows the Ca2+ responses (in false color maps) of a small field of cells exposed to 100 μM glutamate, 1 μM glycine. The 30 sec and 3 min frames show that there was a rapid, but relatively small, increase in intracellular Ca2+ level at 30 sec but almost no subsequent change at the 3 min time point. This was followed by a progressive increase to a plateau over the next 14 min. The population data in Figure 2B obtained from a different experiment encompassing more cells illustrate more clearly the small, rapid onset increase followed by a slowly developing and much larger increase (n = 10). Individual responses measured in the somata of these ten neurons are shown in Figure 2C, demonstrating the similar qualitative features of the increase and also the considerable dispersion in the time courses over which the large delayed increase developed. The fluorescence ratio for Fura-FF in unstimulated cells was ~ 0.12, but this indicator is very insensitive at Ca2+around neuron resting levels owing to its large dissociation constant. Fura-2 measurements in companion cultures showed resting levels to be in the range of 50 –130 nM, which is consistent with measurements from many laboratories (data not shown). The Fura-FF plateau levels reached during glutamate exposure were considerably less than the maximum in situ ratio of 2, produced by ionomycin (see Materials and Methods). Such delayed Ca2+ increases were observed in 60 of the 80 cells obtained from six independent preparations. A similar temporal profile of changes in intracellular Ca2+ was also observed in neurons loaded with Fura-FF and treated with 100 μM NMDA (data not shown).

Figure 2.

Figure 2

Intracellular Ca2+ changes measured with Fura-FF following exposure to glutamate. (A) Initial B&W panel: 380 nm excitation fluorescence picture of the field of cells analyzed. Subsequent panels: false color images showing the fluorescence ratio increases over time following exposure to glutamate. (B) Averaged population data showing the time course of glutamate stimulated intracellular Ca2+ increase measured at neuronal somata (mean ± SEM, n = 10 cells). (C) Individual responses of the 10 neurons are shown to illustrate the typical range of responses observed in different cells. Data are expressed as both a background-corrected fluorescence ratio (left abscissa) and the estimated Ca2+ (right abscissa). Because of the low sensitivity of Fura-FF, Ca2+ estimates below 1 μM are not reliable and should not be extrapolated from the right abscissa.

Although most of the quantifications of Ca2+ changes used measurements made at the soma, we were able to analyze a number of cases where the soma and its main dendrite (out to at least 40 μm were identifiable (n = 26). Relative Ca2+ increases in these two compartments measured at times < 3 min after agonist exposure fell into 3 categories; dendrite > soma (n = 11), dendrite and soma < 30% difference (n = 10), dendrite < soma (n = 5). In all but 4 cases (one illustrated in Fig. 4) the changes had equalized to levels within 30% at times greater than 5 min. It is worth noting that most previous measurements of Ca2+ responses to glutamate have examined much shorter time periods (< 5 min) and have been made using higher affinity indicators. Under these conditions, Ca2+ increases in dendrites can be much larger than those in the somata for the first 1 – 2 min. Indeed we observed this to occur in a proportion of the cells here. However, at the longer time periods we have been interested in here, > 5 min, increases at the soma had “caught up” with dendritic increases. Dendrite measurements are also subject to the consideration that fluorescence is relatively weak and during responses where Ca2+ levels are very high, the 380 nm excited fluorescence (see Materials and Methods) can drop to values close to background, decreasing the reliability of the measurements. Because difference in the measurable Ca2+ values between soma and dendrite are small, and also a large fraction of the total NR2A- and NR2B-NMDA receptors are located at the soma (see Fig. 1), we have employed soma measurements to enable the greatest number of reliable observations. In this context we also note that a previous study (Sinor et al. 2000) has stressed the importance of somatic NR2B receptors in glutamate neurotoxicity.

Figure 4.

Figure 4

Ca2+ recovery enabled by late application of NR2B-NMDAR blocker Ro 25–6981. (A) Fluorescence image of neuron showing measurement locations. (B) Time courses of soma and dendrite Ca2+ changes determined at the locations (see boxes in A). The arrow indicates approximate time of Ro 25–6981 application.

We next examined the effect of NR2B antagonists, Ro 25-6981 and ifenprodil, on glutamate-induced changes in intracellular Ca2+ level. Exposure to Ro 25-6981 before and during glutamate exposure produced only a small decrease in the initial Ca2+ response (4.2 μM to 3.3 μM) measured at the 3 min time point (compare Fig. 2B with 3A). In contrast, the large, late developing phase was blocked (compare Fig. 2B with 3A and 3B). Ifenprodil reduced the initial Ca2+ increases to a greater extent (Fig. 3A), consistent with its less selective blocking profile (Honda et al. 1988; Contreras et al. 1990; Bath et al. 1996). Figure 3C demonstrates that in the presence of glutamate removal of extracellular Ca2+ (no Ca2+ added to media + 2 mM EGTA) brought about recovery from an ongoing glutamate response, arguing that the Ca2+ increases were due to influx and not to Ca2+ release from intracellular stores. MK801, a non-competitive NMDA receptor antagonist, applied before glutamate, blocked nearly all of the Ca2+ increases reported by the low affinity indicator Fura-FF (Fig. 3D). This great reduction occurred despite the activation of AMPA receptors which both depolarize neurons and activate voltage gated Ca2+ channels, and where Ca2+-permeable AMPA receptors are present, directly enable Ca2+ influx (Gorter et al. 1997). Again, it is noted that high affinity indicators readily detect the Ca2+ increases under these conditions.

Figure 3.

Figure 3

NR2B-NMDA receptor-mediated change in intracellular Ca2+ measured with Fura-FF. (A) Population data showing that large, delayed Ca2+ increase does not occur in cells pre-incubated with 1μM Ifenprodil or 10 μM Ro 25–6981 before exposure to glutamate (n = 14). (B) Population data showing that 20 μM Ro 256981 even applied after 5 min exposure to glutamate also prevents the delayed Ca2+ increase (n = 15). (C) Low Ca2+ levels are restored after switch to a Ca2+ free medium still containing 100 μM glutamate (n = 15). (D) The NMDAR channel blocker, MK801, eliminates all changes in the Fura-FF signal (n = 9). Values represent mean ± SEM. Similar results to those shown were obtained from at least 3 independent experiments for all treatments.

The finding that initial Ca2+ increases were very similar in the absence or presence of the selective NR2B antagonist Ro 25-6981 (compare Fig. 2B with 3A) argues that the late Ca2+ increases does not result simply from the loss of regulatory competence of the cells due to the early Ca2+ load. Even late addition of Ro 25-6981 after 5 min incubation with glutamate, prevented the delayed Ca2+ increase (Fig. 3B) with the cells able to regulate the persisting Ca2+ influx for periods of 20 min or more. Moreover the plateau of the late response is well below the saturation level of Fura-FF (see Materials and Methods) indicating that the cells have not simply died. In some instances it was observed that delayed application of the Ro 25-6981 led to an actual decrease in intracellular Ca2+ in the continued presence of glutamate. One such example is given in Figure 4 where both soma and dendrite measurements were possible. Following treatment with glutamate Ca2+ levels displayed their usual increase (1st 5 data sets) till Ro 25-6981 was applied approximately at 5.5 min (as shown by an arrow in Fig. 4B). The subsequent measurements show a clear reduction in intracellular Ca2+ in both the soma and dendrite. Complete recovery is not expected since other glutamate/depolarization pathways should still be active. The above findings imply that in the minutes following glutamate (or NMDA) exposure, an additional influx pathway for Ca2+ mediated by NR2B receptors is recruited which gains importance around 5 min (on average) after the exposure to glutamate or NMDA. This late phase correlates well with both the time course and antagonist sensitivity of STEP dephosphorylation.

The magnitude of intracellular Ca2+ changes regulates STEP dephosphorylation

To determine if the glutamate-mediated dephosphorylation of STEP is dose-dependent, neuronal cultures were exposed to either 10 μM or 100 μM glutamate for 5 and 30 min. Immunoblot analysis of cell lysates with anti-STEP antibody showed that treatment with 10 μM glutamate failed to induce any dephosphorylation of STEP (Fig. 5A, lane 3). However, as shown previously, treatment with 100 μM glutamate for 30 min led to significant dephosphorylation of STEP as evident from the downward shift in the mobility of the STEP band (Fig. 5A, lane 5). We next investigated the dose-dependence of glutamate-mediated Ca2+ increases. Neurons loaded with Fura-FF and exposed to 10 μM glutamate showed a rapid Ca2+ increase of similar magnitude to 100 μM exposure but a much smaller late phase increase (Fig. 5C). The peak Ca2+ level was significantly lower for 10 μM glutamate as compared to neurons exposed to 100 μM glutamate (Fig. 2B).

Figure 5.

Figure 5

Dephosphorylation of STEP is dependent on the magnitude of intracellular changes in Ca2+. Neuron cultures were (A) treated with 10 μM or 100 μM glutamate, (B) treated with 100 μM glutamate or 60 mM KCl, (C) treated with 10 μM glutamate, and (D) treated with 60 mM KCl for the times indicated. (A and B) Equal protein from each sample were analyzed by SDS-PAGE and probed with anti-STEP antibody (upper panel). The blots were re-probed with anti-β tubulin antibody to indicate total protein loading (lower panel). Quantification of phosphorylated (upper band) and non-phosphorylated (lower band) STEP was done using Image J density analysis of immunoblot data obtained from 3 independent experiments. The data is presented as a percentage of the total. Values are mean ± SEM (n = 3). *Indicates significant difference from corresponding control values at 0 min time point (p < 0.001). (C and D) Cells were loaded with Fura-FF prior to treatment with glutamate (n = 13) or KCl (n = 10). Values represent mean ± SEM. Similar results were obtained from at least 3 independent experiments.

In addition to NMDA receptors L-type voltage-gated Ca2+ channels are another principal mediator of Ca2+ entry in neurons and are involved in the activation of ERK2 (Sala et al. 2000; Dolmetsch et al. 2001). But as shown in Figure 5B, and as previously reported KCl-mediated membrane depolarization and influx of Ca2+ failed to dephosphorylate STEP thereby leading to sustained activation of ERK (Paul et al. 2003). In neurons loaded with Fura-FF and exposed to 60 mM KCl there was a relatively small, transient increase in Ca2+, which came back almost to the basal level within 3 min and stayed there for the entire period of the experiment (Fig. 5D). Again the peak Ca2+ level was far lower than that compared to neurons exposed to 100 μM glutamate (Fig. 2B).

DISCUSSION

Our study shows that the intracellular Ca2+ increases occurring after exposure to 100 μM glutamate are very large, requiring the use of a low affinity indicator (here, Fura-FF) to track them, and follow a time course that has two separable phases. In the presence of specific antagonists of NR2B-NMDAR the large delayed increase did not occur, while the smaller, rapid onset phase remained nearly unchanged. In the presence of the non-selective NMDA receptor antagonist MK-801, glutamate application produce no Ca2+ increase measurable with Fura-FF (see Figure 3D), indicating a role for the NR2A-NMDARs in the initial rapid increase in Ca2+. At present there is no selective NR2A-NMDAR antagonist available to provide a better estimate of the Ca2+ increases, due solely to NR2A-NMDAR activation. Zinc has been proposed as a moderately selective and partial blocker for NR2A-NMDAR, but its selectivity is not high enough to make a strong discrimination between the receptor subtypes (Westbrook and Mayer 1987; Chen et al. 1997a; Paoletti et al. 1997; Choi and Lipton 1999; Madry et al. 2007). Evidences also suggest that zinc exposure might up-regulate NMDA receptor activity via activation of Src family kinases (Manzerra et al. 2001; Kim et al. 2002). Besides modulating NMDA receptor function zinc also acts on other post-synaptic channels and receptors including voltage gated ion channels, GABA (γ-aminobutyric acid), dopamine and Trk (tyrosine kinase) receptors (Bitanihirwe and Cunningham 2009) and therefore cannot be used as a selective inhibitor for NR2A-NMDA receptors. More recent studies reported the discovery of a new NMDAR antaogonist, NVP-AAM077 ([(R)-[(S)-1-(4-bromophenyl)-ethylamino]-(2, 3-dioxo-1,2,3,4-tetrahydroquinoxalin-5-yl)-methyl]-phosphonic acid)) (Auberson et al. 2002), which was claimed to display strong selectivity for NR2A-containing NMDA receptors (Liu et al. 2004). However, its selectivity has been debated in several studies (Feng et al. 2004; Berberich et al. 2005; Weitlauf et al. 2005; Frizelle et al. 2006; Neyton and Paoletti 2006; de Marchena et al. 2008). It is also not commercially available.

The rapid onset, but relatively low level Ca2+ increases which we hypothesize here are largely due to NR2A-NMDAR activation, are sufficient to activate ERK2 but had no effect on the tyrosine phosphatase STEP, allowing ERK to remain active throughout the time course examined. On the other hand, the delayed larger Ca2+ increase resulting from NR2B-NMDAR activation leads to the dephosphorylation and subsequent activation of STEP. As a consequence, activated STEP dephosphorylates ERK thereby limiting the duration of its activation.

The two phase nature of Ca2+ increases following exposures to high levels of glutamate can be seen in published work but the phases are much less distinct than we report here and have not been remarked upon (Hyrc et al. 1997; Cheng et al. 1999). A particular contribution of our study is the identification of the late increase being due to NR2B-NMDAR's and not to other factors such as failure of Ca2+ regulatory competence. This is not to say that such failure does not occur at later times. The blocker Ro 25-6981 had a very small effect on the early phase of Ca2+ increase from which we hypothesize that this initial increase is largely due to NR2A-NMDAR's and voltage-gated channels. The small size of the early Ca2+ increase relative to the later NR2B-NMDAR mediated increase is at least consistent with the developmental stage of the cultures used (~ DIV 14). Cheng et al., (1999) have shown that NR2A expression is < 50% of that reached at later times, while NR2B is ~ 150% of later values (Stocca and Vicini 1998). There is a well characterized, Ca2+ dependent inactivation of the current carried by NMDA receptors (Krupp et al. 1996; Kyrozis et al. 1996), but the relatively small size of the presumed NR2A dependent Ca2+ increase phase cannot be attributed to this phenomenon, since this inactivation is very rapid, occurring over a period of a few seconds and would be missed in the slow, low affinity measurements used here. The reason for the delay in onset of the NR2B mediated response is not known. It certainly cannot be due to channel kinetics since the time scale is minutes. We propose that it reflects the recruitment of NR2B-NMDAR's into a functional state from a pool of non-functional receptors and that investigation of the mechanisms that bring this about is an important topic for future study.

We have regarded the primary underlying cause of the Ca2+ increases seen here to be trans-membrane influx of Ca2+ rather than the other well studied cause, Ca2+ release from intra-cellular stores. Although receptor activated Ca2+ increases, due to intracellular release, have been demonstrated in cortical culture preparation as well as in other neuronal preparations (Pozzo-Miller et al. 2000) these increases have been much smaller, in the sub micromolar range, and are transient, lasting a few seconds at most. Changes reported here range from 2–30 μM upward and persist for many minutes, so that while intracellular release may possibly underlie a part of the responses, the contribution is small. Moreover the decrease in intracellular Ca2+ level a. lmost to base line following removal of extracellular Ca2+ 5 min after glutamate treatment further indicates that the underlying cause of these Ca2+ increases to be trans-membrane influx of Ca2+. Consistent with this interpretation, electrophysiological studies from other laboratories have shown a similar biphasic change in cationic currents in acutely dissociated hippocampal neurons following exposure to 100 μM NMDA (Chen et al. 1997b; Thompson et al. 2008). The Chen et al., (1997) study was made before NR2B antagonists were available. The Thompson et al., (2008) study demonstrated the response in hippocampal cultures also. Here, there was a rapidly developing inward current, carried partially by Ca2+ that would give rise to the rapid Ca2+ increase, followed by a larger, slowly developing current that would give rise to the larger, more slowly developing Ca2+ increase. In the study of Thompson et al., the slow developing cation current was attributed to the opening of Pannexin-1 hemichannels in the plasma membrane, this late developing current being blocked by either carbenoxolone, or a small peptide inhibitor, or RNAi of Pannexin1. In our study, the NR2B antagonists, ifenprodil and Ro 25-6981 blocked the delayed Ca2+ increase but carbenoxolone did not (data not shown). There was also no significant loss of indicator in our studies as occurs with hemichannel opening (Thompson et al. 2008). The straightforward interpretation of the findings is that different mechanisms are operative in different subtype of neurons.

Primary cortical neuronal cultures treated with high concentrations of glutamate (50 –100 μM) are commonly used as a model of neuronal injury. In severe brain disorders such as ischemic stroke and traumatic brain injury a marked increase in the release of glutamate has also been observed and is responsible for NMDAR-mediated brain damage (Simon et al. 1984; Choi and Rothman 1990; Michaels and Rothman 1990; Globus et al. 1995; Rossi et al. 2000). Direct measurement of extracellular glutamate concentration with in vivo microdialysis has shown that experimental ischemia for more than 30 min causes extracellular glutamate levels to rise from under 5 μM to more than 500 μM (Benveniste et al. 1984; Hagberg et al. 1985; Choi and Rothman 1990). Although the origin of the increased extracellular glutamate has not been unequivocally established it has been suggested that reversed operation of both neuronal and glial transporters contribute to the excessive glutamate burden (Rossi et al. 2000). Few studies have also addressed the role of NR2A and NR2B receptor subtypes in glutamate mediated excitotoxicity in vivo (Morikawa et al. 1998; Liu et al. 2007; Tu et al. 2010). The study by Morikawa and colleagues (1998) reported that generating ischemic stroke in mutant mice that lack NR2A resulted in significant reduction in brain damage suggesting that NR2A plays a role in glutamate neurotoxicity. More recent studies, using animal models of ischemic stroke, have reported a role of NR2B-NMDAR in ischemic brain damage (Liu et al. 2007; Tu et al. 2010). The study by Tu et al., (2010) shows that interaction of DAPK1 (death associated protein kinase 1) with NR2B-NMDAR complex mediates brain damage in ischemic stroke and disruption of this association provides neuroprotection. These latter findings imply NR2B-NMDARs are also an important mediator of excitotoxicity not only in neuronal cultures but in mature neurons in vivo as well. Knowledge regarding the specific intracellular signaling pathways leading to glutamate / NMDAR-induced toxicity is incomplete. However, several in vivo studies have shown that ERK1/2 plays a crucial role in ischemic brain injury and specific inhibitors of ERK signaling pathway provides protection against forebrain ischemia (Alessandrini et al. 1999; Irving et al. 2000; Namura et al. 2001). In this context we report that relatively prolonged (> 5 min) treatment with an excitotoxic dose of glutamate (100 μM) leads to NR2B-NMDAR mediated dephosphorylation and subsequent activation of STEP. Since STEP is involved in the inactivation of ERK, the findings suggest that STEP can play an initial role in neuronal cell survival by preventing sustained activation of ERK. However, if the extracellular insult is sustained this protection may be ineffective, since extended exposure to an excitotoxic dose of glutamate may lead to breakdown of STEP and subsequent activation other degenerative processes. Consistent with this hypothesis, a recent study using an animal model of ischemic stroke has shown that STEP is unilaterally down regulated in the stroked hemisphere over time. Furthermore, STEP levels are increased in a neuroprotective-preconditioning paradigm in regions of the brain that are resistant to ischemic brain damage (Braithwaite et al. 2008). Taken together these findings suggest that STEP may be considered as a significant neuroprotectant against brief excitatory insults.

ACKNOWLEDGEMENTS

This work was supported by the National Institutes of Health grants NS059962 (Paul, S) and P20 RR015636 (Okada, Y). We would like to thank Drs. Ranjana Poddar and Ishani Deb for technical assistance and Dr. Angus C Nairn, Yale University, for his helpful comments.

Abbreviations

ERK

extracellular-regulated kinase

MAP

mitogen-activated protein

APV

DL-2-amino-5-phosphonopentanoic acid

CNQX

6-cyano-7-nitroquinoxaline-2,3-dione

STEP

striatal-enriched phosphatase

SDS

sodium dodecyl sulfate

PKA

protein kinase A

DIV

days in vitro

REFERENCE

  1. Alessandrini A, Namura S, Moskowitz MA, Bonventre JV. MEK1 protein kinase inhibition protects against damage resulting from focal cerebral ischemia. Proceedings of the National Academy of Sciences of the United States of America. 1999;96:12866–12869. doi: 10.1073/pnas.96.22.12866. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Auberson YP, Allgeier H, Bischoff S, Lingenhoehl K, Moretti R, Schmutz M. 5-Phosphonomethylquinoxalinediones as competitive NMDA receptor antagonists with a preference for the human 1A/2A, rather than 1A/2B receptor composition. Bioorganic & medicinal chemistry letters. 2002;12:1099–1102. doi: 10.1016/s0960-894x(02)00074-4. [DOI] [PubMed] [Google Scholar]
  3. Banko JL, Hou L, Klann E. NMDA receptor activation results in PKA- and ERK-dependent Mnk1 activation and increased eIF4E phosphorylation in hippocampal area CA1. Journal of neurochemistry. 2004;91:462–470. doi: 10.1111/j.1471-4159.2004.02734.x. [DOI] [PubMed] [Google Scholar]
  4. Bath CP, Farrell LN, Gilmore J, Ward MA, Hicks CA, O'Neill MJ, Bleakman D. The effects of ifenprodil and eliprodil on voltage-dependent Ca2+ channels and in gerbil global cerebral ischaemia. European journal of pharmacology. 1996;299:103–112. doi: 10.1016/0014-2999(95)00846-2. [DOI] [PubMed] [Google Scholar]
  5. Benveniste H, Drejer J, Schousboe A, Diemer NH. Elevation of the extracellular concentrations of glutamate and aspartate in rat hippocampus during transient cerebral ischemia monitored by intracerebral microdialysis. Journal of neurochemistry. 1984;43:1369–1374. doi: 10.1111/j.1471-4159.1984.tb05396.x. [DOI] [PubMed] [Google Scholar]
  6. Berberich S, Punnakkal P, Jensen V, Pawlak V, Seeburg PH, Hvalby O, Kohr G. Lack of NMDA receptor subtype selectivity for hippocampal long-term potentiation. J Neurosci. 2005;25:6907–6910. doi: 10.1523/JNEUROSCI.1905-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bitanihirwe BK, Cunningham MG. Zinc: the brain's dark horse. Synapse (New York, N.Y. 2009;63:1029–1049. doi: 10.1002/syn.20683. [DOI] [PubMed] [Google Scholar]
  8. Boulanger LM, Lombroso PJ, Raghunathan A, During MJ, Wahle P, Naegele JR. Cellular and molecular characterization of a brain-enriched protein tyrosine phosphatase. J Neurosci. 1995;15:1532–1544. doi: 10.1523/JNEUROSCI.15-02-01532.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Braithwaite SP, Adkisson M, Leung J, Nava A, Masterson B, Urfer R, Oksenberg D, Nikolich K. Regulation of NMDA receptor trafficking and function by striatal-enriched tyrosine phosphatase (STEP) The European journal of neuroscience. 2006;23:2847–2856. doi: 10.1111/j.1460-9568.2006.04837.x. [DOI] [PubMed] [Google Scholar]
  10. Braithwaite SP, Xu J, Leung J, Urfer R, Nikolich K, Oksenberg D, Lombroso PJ, Shamloo M. Expression and function of striatal enriched protein tyrosine phosphatase is profoundly altered in cerebral ischemia. The European journal of neuroscience. 2008;27:2444–2452. doi: 10.1111/j.1460-9568.2008.06209.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Bult A, Zhao F, Dirkx R, Jr., Raghunathan A, Solimena M, Lombroso PJ. STEP: a family of brain-enriched PTPs. Alternative splicing produces transmembrane, cytosolic and truncated isoforms. European journal of cell biology. 1997;72:337–344. [PubMed] [Google Scholar]
  12. Chandler LJ, Sutton G, Dorairaj NR, Norwood D. N-methyl D-aspartate receptor-mediated bidirectional control of extracellular signal-regulated kinase activity in cortical neuronal cultures. The Journal of biological chemistry. 2001;276:2627–2636. doi: 10.1074/jbc.M003390200. [DOI] [PubMed] [Google Scholar]
  13. Charton JP, Herkert M, Becker CM, Schroder H. Cellular and subcellular localization of the 2B-subunit of the NMDA receptor in the adult rat telencephalon. Brain Res. 1999;816:609–617. doi: 10.1016/s0006-8993(98)01243-8. [DOI] [PubMed] [Google Scholar]
  14. Chen N, Moshaver A, Raymond LA. Differential sensitivity of recombinant N-methyl-D-aspartate receptor subtypes to zinc inhibition. Molecular pharmacology. 1997a;51:1015–1023. doi: 10.1124/mol.51.6.1015. [DOI] [PubMed] [Google Scholar]
  15. Chen QX, Perkins KL, Choi DW, Wong RK. Secondary activation of a cation conductance is responsible for NMDA toxicity in acutely isolated hippocampal neurons. J Neurosci. 1997b;17:4032–4036. doi: 10.1523/JNEUROSCI.17-11-04032.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Cheng C, Fass DM, Reynolds IJ. Emergence of excitotoxicity in cultured forebrain neurons coincides with larger glutamate-stimulated [Ca(2+)](i) increases and NMDA receptor mRNA levels. Brain Res. 1999;849:97–108. doi: 10.1016/s0006-8993(99)01995-2. [DOI] [PubMed] [Google Scholar]
  17. Choi DW, Rothman SM. The role of glutamate neurotoxicity in hypoxic-ischemic neuronal death. Annual review of neuroscience. 1990;13:171–182. doi: 10.1146/annurev.ne.13.030190.001131. [DOI] [PubMed] [Google Scholar]
  18. Choi YB, Lipton SA. Identification and mechanism of action of two histidine residues underlying high-affinity Zn2+ inhibition of the NMDA receptor. Neuron. 1999;23:171–180. doi: 10.1016/s0896-6273(00)80763-1. [DOI] [PubMed] [Google Scholar]
  19. Connor JA. Digital imaging of free calcium changes and of spatial gradients in growing processes in single, mammalian central nervous system cells. Proceedings of the National Academy of Sciences of the United States of America. 1986;83:6179–6183. doi: 10.1073/pnas.83.16.6179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Contreras PC, Bremer ME, Gray NM. Ifenprodil and SL 82.0715 potently inhibit binding of [3H](+)-3-PPP to sigma binding sites in rat brain. Neuroscience letters. 1990;116:190–193. doi: 10.1016/0304-3940(90)90408-2. [DOI] [PubMed] [Google Scholar]
  21. Crozier RA, Black IB, Plummer MR. Blockade of NR2B-containing NMDA receptors prevents BDNF enhancement of glutamatergic transmission in hippocampal neurons. Learning & memory (Cold Spring Harbor, N.Y. 1999;6:257–266. [PMC free article] [PubMed] [Google Scholar]
  22. de Marchena J, Roberts AC, Middlebrooks PG, Valakh V, Yashiro K, Wilfley LR, Philpot BD. NMDA receptor antagonists reveal age-dependent differences in the properties of visual cortical plasticity. Journal of neurophysiology. 2008;100:1936–1948. doi: 10.1152/jn.90290.2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Dolmetsch RE, Pajvani U, Fife K, Spotts JM, Greenberg ME. Signaling to the nucleus by an L-type calcium channel-calmodulin complex through the MAP kinase pathway. Science (New York, N.Y. 2001;294:333–339. doi: 10.1126/science.1063395. [DOI] [PubMed] [Google Scholar]
  24. Feng B, Tse HW, Skifter DA, Morley R, Jane DE, Monaghan DT. Structure-activity analysis of a novel NR2C/NR2D-preferring NMDA receptor antagonist: 1-(phenanthrene-2-carbonyl) piperazine-2,3-dicarboxylic acid. British journal of pharmacology. 2004;141:508–516. doi: 10.1038/sj.bjp.0705644. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Fischer G, Mutel V, Trube G, Malherbe P, Kew JN, Mohacsi E, Heitz MP, Kemp JA. Ro 25-6981, a highly potent and selective blocker of N-methyl-D-aspartate receptors containing the NR2B subunit. Characterization in vitro. The Journal of pharmacology and experimental therapeutics. 1997;283:1285–1292. [PubMed] [Google Scholar]
  26. Frizelle PA, Chen PE, Wyllie DJ. Equilibrium constants for (R)-[(S)-1-(4-bromo-phenyl)-ethylamino]-(2,3-dioxo-1,2,3,4-tetrahydroquino xalin-5-yl)-methyl]-phosphonic acid (NVP-AAM077) acting at recombinant NR1/NR2A and NR1/NR2B N-methyl-D-aspartate receptors: Implications for studies of synaptic transmission. Molecular pharmacology. 2006;70:1022–1032. doi: 10.1124/mol.106.024042. [DOI] [PubMed] [Google Scholar]
  27. Globus MY, Alonso O, Dietrich WD, Busto R, Ginsberg MD. Glutamate release and free radical production following brain injury: effects of posttraumatic hypothermia. Journal of neurochemistry. 1995;65:1704–1711. doi: 10.1046/j.1471-4159.1995.65041704.x. [DOI] [PubMed] [Google Scholar]
  28. Gorter JA, Petrozzino JJ, Aronica EM, Rosenbaum DM, Opitz T, Bennett MV, Connor JA, Zukin RS. Global ischemia induces downregulation of Glur2 mRNA and increases AMPA receptor-mediated Ca2+ influx in hippocampal CA1 neurons of gerbil. J Neurosci. 1997;17:6179–6188. doi: 10.1523/JNEUROSCI.17-16-06179.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Grynkiewicz G, Poenie M, Tsien RY. A new generation of Ca2+ indicators with greatly improved fluorescence properties. The Journal of biological chemistry. 1985;260:3440–3450. [PubMed] [Google Scholar]
  30. Hagberg H, Lehmann A, Sandberg M, Nystrom B, Jacobson I, Hamberger A. Ischemia-induced shift of inhibitory and excitatory amino acids from intra- to extracellular compartments. J Cereb Blood Flow Metab. 1985;5:413–419. doi: 10.1038/jcbfm.1985.56. [DOI] [PubMed] [Google Scholar]
  31. Hallett PJ, Spoelgen R, Hyman BT, Standaert DG, Dunah AW. Dopamine D1 activation potentiates striatal NMDA receptors by tyrosine phosphorylation-dependent subunit trafficking. J Neurosci. 2006;26:4690–4700. doi: 10.1523/JNEUROSCI.0792-06.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Honda H, Sakai Y, Iwata T, Ohba S, Kanezuka T, Irino O. Effects of ifenprodil tartrate on alpha-adrenoceptors and Ca2+ movement in isolated canine saphenous veins. Archives internationales de pharmacodynamie et de therapie. 1988;292:112–121. [PubMed] [Google Scholar]
  33. Hyrc K, Handran SD, Rothman SM, Goldberg MP. Ionized intracellular calcium concentration predicts excitotoxic neuronal death: observations with low-affinity fluorescent calcium indicators. J Neurosci. 1997;17:6669–6677. doi: 10.1523/JNEUROSCI.17-17-06669.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Irving EA, Barone FC, Reith AD, Hadingham SJ, Parsons AA. Differential activation of MAPK/ERK and p38/SAPK in neurones and glia following focal cerebral ischaemia in the rat. Brain research. 2000;77:65–75. doi: 10.1016/s0169-328x(00)00043-7. [DOI] [PubMed] [Google Scholar]
  35. Jiang Q, Gu Z, Zhang G, Jing G. Diphosphorylation and involvement of extracellular signal-regulated kinases (ERK1/2) in glutamate-induced apoptotic-like death in cultured rat cortical neurons. Brain Res. 2000a;857:71–77. doi: 10.1016/s0006-8993(99)02364-1. [DOI] [PubMed] [Google Scholar]
  36. Jiang Q, Gu Z, Zhang G, Jing G. N-methyl-D-aspartate receptor activation results in regulation of extracellular signal-regulated kinases by protein kinases and phosphatases in glutamate-induced neuronal apototic-like death. Brain Res. 2000b;887:285–292. doi: 10.1016/s0006-8993(00)03003-1. [DOI] [PubMed] [Google Scholar]
  37. Kaphzan H, O'Riordan KJ, Mangan KP, Levenson JM, Rosenblum K. NMDA and dopamine converge on the NMDA-receptor to induce ERK activation and synaptic depression in mature hippocampus. PLoS ONE. 2006;1:e138. doi: 10.1371/journal.pone.0000138. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Kim MJ, Dunah AW, Wang YT, Sheng M. Differential roles of NR2A- and NR2B-containing NMDA receptors in Ras-ERK signaling and AMPA receptor trafficking. Neuron. 2005;46:745–760. doi: 10.1016/j.neuron.2005.04.031. [DOI] [PubMed] [Google Scholar]
  39. Kim TY, Hwang JJ, Yun SH, Jung MW, Koh JY. Augmentation by zinc of NMDA receptor-mediated synaptic responses in CA1 of rat hippocampal slices: mediation by Src family tyrosine kinases. Synapse (New York, N.Y. 2002;46:49–56. doi: 10.1002/syn.10118. [DOI] [PubMed] [Google Scholar]
  40. Kohr G. NMDA receptor function: subunit composition versus spatial distribution. Cell and tissue research. 2006;326:439–446. doi: 10.1007/s00441-006-0273-6. [DOI] [PubMed] [Google Scholar]
  41. Komiyama NH, Watabe AM, Carlisle HJ, Porter K, Charlesworth P, Monti J, Strathdee DJ, O'Carroll CM, Martin SJ, Morris RG, O'Dell TJ, Grant SG. SynGAP regulates ERK/MAPK signaling, synaptic plasticity, and learning in the complex with postsynaptic density 95 and NMDA receptor. J Neurosci. 2002;22:9721–9732. doi: 10.1523/JNEUROSCI.22-22-09721.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Krupp JJ, Vissel B, Heinemann SF, Westbrook GL. Calcium-dependent inactivation of recombinant N-methyl-D-aspartate receptors is NR2 subunit specific. Molecular pharmacology. 1996;50:1680–1688. [PubMed] [Google Scholar]
  43. Kyrozis A, Albuquerque C, Gu J, MacDermott AB. Ca(2+)-dependent inactivation of NMDA receptors: fast kinetics and high Ca2+ sensitivity in rat dorsal horn neurons. The Journal of physiology. 1996;495(Pt 2):449–463. doi: 10.1113/jphysiol.1996.sp021606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  44. Liu L, Wong TP, Pozza MF, Lingenhoehl K, Wang Y, Sheng M, Auberson YP, Wang YT. Role of NMDA receptor subtypes in governing the direction of hippocampal synaptic plasticity. Science (New York, N.Y. 2004;304:1021–1024. doi: 10.1126/science.1096615. [DOI] [PubMed] [Google Scholar]
  45. Liu Y, Wong TP, Aarts M, Rooyakkers A, Liu L, Lai TW, Wu DC, Lu J, Tymianski M, Craig AM, Wang YT. NMDA receptor subunits have differential roles in mediating excitotoxic neuronal death both in vitro and in vivo. J Neurosci. 2007;27:2846–2857. doi: 10.1523/JNEUROSCI.0116-07.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Lombroso PJ, Naegele JR, Sharma E, Lerner M. A protein tyrosine phosphatase expressed within dopaminoceptive neurons of the basal ganglia and related structures. J Neurosci. 1993;13:3064–3074. doi: 10.1523/JNEUROSCI.13-07-03064.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Madry C, Mesic I, Betz H, Laube B. The N-terminal domains of both NR1 and NR2 subunits determine allosteric Zn2+ inhibition and glycine affinity of N-methyl-D-aspartate receptors. Molecular pharmacology. 2007;72:1535–1544. doi: 10.1124/mol.107.040071. [DOI] [PubMed] [Google Scholar]
  48. Manzerra P, Behrens MM, Canzoniero LM, Wang XQ, Heidinger V, Ichinose T, Yu SP, Choi DW. Zinc induces a Src family kinase-mediated up-regulation of NMDA receptor activity and excitotoxicity. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:11055–11061. doi: 10.1073/pnas.191353598. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Martel MA, Wyllie DJ, Hardingham GE. In developing hippocampal neurons, NR2B-containing N-methyl-D-aspartate receptors (NMDARs) can mediate signaling to neuronal survival and synaptic potentiation, as well as neuronal death. Neuroscience. 2009;158:334–343. doi: 10.1016/j.neuroscience.2008.01.080. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Michaels RL, Rothman SM. Glutamate neurotoxicity in vitro: antagonist pharmacology and intracellular calcium concentrations. J Neurosci. 1990;10:283–292. doi: 10.1523/JNEUROSCI.10-01-00283.1990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Morikawa E, Mori H, Kiyama Y, Mishina M, Asano T, Kirino T. Attenuation of focal ischemic brain injury in mice deficient in the epsilon1 (NR2A) subunit of NMDA receptor. J Neurosci. 1998;18:9727–9732. doi: 10.1523/JNEUROSCI.18-23-09727.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Namura S, Iihara K, Takami S, Nagata I, Kikuchi H, Matsushita K, Moskowitz MA, Bonventre JV, Alessandrini A. Intravenous administration of MEK inhibitor U0126 affords brain protection against forebrain ischemia and focal cerebral ischemia. Proceedings of the National Academy of Sciences of the United States of America. 2001;98:11569–11574. doi: 10.1073/pnas.181213498. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Neyton J, Paoletti P. Relating NMDA receptor function to receptor subunit composition: limitations of the pharmacological approach. J Neurosci. 2006;26:1331–1333. doi: 10.1523/JNEUROSCI.5242-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Paoletti P, Ascher P, Neyton J. High-affinity zinc inhibition of NMDA NR1-NR2A receptors. J Neurosci. 1997;17:5711–5725. doi: 10.1523/JNEUROSCI.17-15-05711.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Paquet M, Tremblay M, Soghomonian JJ, Smith Y. AMPA and NMDA glutamate receptor subunits in midbrain dopaminergic neurons in the squirrel monkey: an immunohistochemical and in situ hybridization study. J Neurosci. 1997;17:1377–1396. doi: 10.1523/JNEUROSCI.17-04-01377.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Paul S, Nairn AC, Wang P, Lombroso PJ. NMDA-mediated activation of the tyrosine phosphatase STEP regulates the duration of ERK signaling. Nature neuroscience. 2003;6:34–42. doi: 10.1038/nn989. [DOI] [PubMed] [Google Scholar]
  57. Paul S, Snyder GL, Yokakura H, Picciotto MR, Nairn AC, Lombroso PJ. The Dopamine/D1 receptor mediates the phosphorylation and inactivation of the protein tyrosine phosphatase STEP via a PKA-dependent pathway. J Neurosci. 2000;20:5630–5638. doi: 10.1523/JNEUROSCI.20-15-05630.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  58. Paul S, Olausson P, Venkitaramani DV, Ruchkina I, Moran TD, Tronson N, Mills E, Hakim S, Salter MW, Taylor JR, Lombroso PJ. The striatal-enriched protein tyrosine phosphatase gates long-term potentiation and fear memory in the lateral amygdala. Biological psychiatry. 2007;61:1049–1061. doi: 10.1016/j.biopsych.2006.08.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Pelkey KA, Askalan R, Paul S, Kalia LV, Nguyen TH, Pitcher GM, Salter MW, Lombroso PJ. Tyrosine phosphatase STEP is a tonic brake on induction of long-term potentiation. Neuron. 2002;34:127–138. doi: 10.1016/s0896-6273(02)00633-5. [DOI] [PubMed] [Google Scholar]
  60. Petralia RS, Wang YX, Wenthold RJ. The NMDA receptor subunits NR2A and NR2B show histological and ultrastructural localization patterns similar to those of NR1. J Neurosci. 1994;14:6102–6120. doi: 10.1523/JNEUROSCI.14-10-06102.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Petrozzino JJ, Pozzo Miller LD, Connor JA. Micromolar Ca2+ transients in dendritic spines of hippocampal pyramidal neurons in brain slice. Neuron. 1995;14:1223–1231. doi: 10.1016/0896-6273(95)90269-4. [DOI] [PubMed] [Google Scholar]
  62. Pozzo-Miller LD, Connor JA, Andrews SB. Microheterogeneity of calcium signalling in dendrites. The Journal of physiology. 2000;525(Pt 1):53–61. doi: 10.1111/j.1469-7793.2000.t01-1-00053.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Rao A, Kim E, Sheng M, Craig AM. Heterogeneity in the molecular composition of excitatory postsynaptic sites during development of hippocampal neurons in culture. J Neurosci. 1998;18:1217–1229. doi: 10.1523/JNEUROSCI.18-04-01217.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Rossi DJ, Oshima T, Attwell D. Glutamate release in severe brain ischaemia is mainly by reversed uptake. Nature. 2000;403:316–321. doi: 10.1038/35002090. [DOI] [PubMed] [Google Scholar]
  65. Sala C, Rudolph-Correia S, Sheng M. Developmentally regulated NMDA receptor-dependent dephosphorylation of cAMP response element-binding protein (CREB) in hippocampal neurons. J Neurosci. 2000;20:3529–3536. doi: 10.1523/JNEUROSCI.20-10-03529.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  66. Sharma E, Zhao F, Bult A, Lombroso PJ. Identification of two alternatively spliced transcripts of STEP: a subfamily of brain-enriched protein tyrosine phosphatases. Brain research. 1995;32:87–93. doi: 10.1016/0169-328x(95)00066-2. [DOI] [PubMed] [Google Scholar]
  67. Simon RP, Swan JH, Griffiths T, Meldrum BS. Blockade of N-methyl-D-aspartate receptors may protect against ischemic damage in the brain. Science (New York, N.Y. 1984;226:850–852. doi: 10.1126/science.6093256. [DOI] [PubMed] [Google Scholar]
  68. Sinor JD, Du S, Venneti S, Blitzblau RC, Leszkiewicz DN, Rosenberg PA, Aizenman E. NMDA and glutamate evoke excitotoxicity at distinct cellular locations in rat cortical neurons in vitro. J Neurosci. 2000;20:8831–8837. doi: 10.1523/JNEUROSCI.20-23-08831.2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  69. Stocca G, Vicini S. Increased contribution of NR2A subunit to synaptic NMDA receptors in developing rat cortical neurons. The Journal of physiology. 1998;507(Pt 1):13–24. doi: 10.1111/j.1469-7793.1998.013bu.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  70. Stout AK, Reynolds IJ. High-affinity calcium indicators underestimate increases in intracellular calcium concentrations associated with excitotoxic glutamate stimulations. Neuroscience. 1999;89:91–100. doi: 10.1016/s0306-4522(98)00441-2. [DOI] [PubMed] [Google Scholar]
  71. Thomas GM, Huganir RL. MAPK cascade signalling and synaptic plasticity. Nature reviews. 2004;5:173–183. doi: 10.1038/nrn1346. [DOI] [PubMed] [Google Scholar]
  72. Thompson RJ, Jackson MF, Olah ME, Rungta RL, Hines DJ, Beazely MA, MacDonald JF, MacVicar BA. Activation of pannexin-1 hemichannels augments aberrant bursting in the hippocampus. Science (New York, N.Y. 2008;322:1555–1559. doi: 10.1126/science.1165209. [DOI] [PubMed] [Google Scholar]
  73. Tongiorgi E, Cattaneo A, Domenici L. Co-expression of TrkB and the N-methyl-D-aspartate receptor subunits NR1-C1, NR2A and NR2B in the rat visual cortex. Neuroscience. 1999;90:1361–1369. doi: 10.1016/s0306-4522(98)00473-4. [DOI] [PubMed] [Google Scholar]
  74. Tovar KR, Westbrook GL. The incorporation of NMDA receptors with a distinct subunit composition at nascent hippocampal synapses in vitro. J Neurosci. 1999;19:4180–4188. doi: 10.1523/JNEUROSCI.19-10-04180.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  75. Tu W, Xu X, Peng L, Zhong X, Zhang W, Soundarapandian MM, Balel C, Wang M, Jia N, Zhang W, Lew F, Chan SL, Chen Y, Lu Y. DAPK1 interaction with NMDA receptor NR2B subunits mediates brain damage in stroke. Cell. 2010;140:222–234. doi: 10.1016/j.cell.2009.12.055. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Valjent E, Pascoli V, Svenningsson P, Paul S, Enslen H, Corvol JC, Stipanovich A, Caboche J, Lombroso PJ, Nairn AC, Greengard P, Herve D, Girault JA. Regulation of a protein phosphatase cascade allows convergent dopamine and glutamate signals to activate ERK in the striatum. Proceedings of the National Academy of Sciences of the United States of America. 2005;102:491–496. doi: 10.1073/pnas.0408305102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Waxman EA, Lynch DR. N-methyl-D-aspartate receptor subtype mediated bidirectional control of p38 mitogen-activated protein kinase. The Journal of biological chemistry. 2005;280:29322–29333. doi: 10.1074/jbc.M502080200. [DOI] [PubMed] [Google Scholar]
  78. Weitlauf C, Honse Y, Auberson YP, Mishina M, Lovinger DM, Winder DG. Activation of NR2A-containing NMDA receptors is not obligatory for NMDA receptor-dependent long-term potentiation. J Neurosci. 2005;25:8386–8390. doi: 10.1523/JNEUROSCI.2388-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  79. Westbrook GL, Mayer ML. Micromolar concentrations of Zn2+ antagonize NMDA and GABA responses of hippocampal neurons. Nature. 1987;328:640–643. doi: 10.1038/328640a0. [DOI] [PubMed] [Google Scholar]
  80. Williams K. Ifenprodil, a novel NMDA receptor antagonist: site and mechanism of action. Current drug targets. 2001;2:285–298. doi: 10.2174/1389450013348489. [DOI] [PubMed] [Google Scholar]
  81. Yang J, Woodhall GL, Jones RS. Tonic facilitation of glutamate release by presynaptic NR2B-containing NMDA receptors is increased in the entorhinal cortex of chronically epileptic rats. J Neurosci. 2006;26:406–410. doi: 10.1523/JNEUROSCI.4413-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES