pmc logo imageJournal ListSearchpmc logo image
Logo of nihpaNIHPA bannerabout author manuscriptssubmit a manuscript
Neuroscience.Author manuscript; available in PMC 2006 June 5.
Published in final edited form as:
PMCID: PMC1473990
NIHMSID: NIHMS9621
CALCINEURIN ENHANCES L-TYPE Ca2+ CHANNEL ACTIVITY IN HIPPOCAMPAL NEURONS: INCREASED EFFECT WITH AGE IN CULTURE
C. M. NORRIS,* E. M. BLALOCK, K.-C. CHEN, N. M. PORTER, and P. W. LANDFIELD
Department of Molecular and Biomedical Pharmacology, College of Medicine, MS-310 UKMC, University of Kentucky, Lexington, KY 40536-0298, USA
*Corresponding author. Tel.: +1-859-323-6124; fax: +1-859-323-1981. E-mail address:cnorr2/at/pop.uky.edu (C. M. Norris).
Abstract
The Ca2+/calmodulin-dependent protein phosphatase, calcineurin, modulates a number of key Ca2+ signaling pathways in neurons, and has been implicated in Ca2+-dependent negative feedback inactivation of N-methyl-d-aspartate receptors and voltage-sensitive Ca2+ channels. In contrast, we report here that three mechanistically disparate calcineurin inhibitors, FK-506, cyclosporin A, and the calcineurin autoinhibitory peptide, inhibited high-voltage-activated Ca2+ channel currents by up to 40% in cultured hippocampal neurons, suggesting that calcineurin acts to enhance Ca2+ currents. This effect occurred with Ba2+ or Ca2+ as charge carrier, and with or without intracellular Ca2+ buffered by EGTA. Ca2+-dependent inactivation of Ca2+ channels was not affected by FK-506. The immunosuppressant, rapamycin, and the protein phosphatase 1/2A inhibitor, okadaic acid, did not decrease Ca2+ channel current, showing specificity for effects on calcineurin. Blockade of L-type Ca2+ channels with nimodipine fully negated the effect of FK-506 on Ca2+ channel current, while blockade of N-, and P-/Q-type Ca2+ channels enhanced FK-506-mediated inhibition of the remaining L-type-enriched current. FK-506 also inhibited substantially more Ca2+ channel current in 4-week-old vs. 2-week-old cultures, an effect paralleled by an increase in calcineurin A mRNA levels.

These studies provide the first evidence that calcineurin selectively enhances L-type Ca2+ channel activity in neurons. Moreover, this action appears to be increased concomitantly with the well-characterized increase in L-type Ca2+ channel availability in hippocampal neurons with age-in-culture.

Keywords: protein phosphatase, aging, Ca2+ channel currents, FK-506, cyclosporin A, nimodipine, conotoxins
Abbreviations: ANOVA, analysis of variance; [Ca2+]i,intracellular Ca2+ concentration; CN-AIP, calcineurin autoinhibitory peptide; CNQX, 6-cyano-7-nitroquinoxaline-2,3-dione disodium; CsA, cyclosporin A; DIV, days in vitro; EDTA, ethylene glycol-bis(2-aminoethyl-ether)-N,N,N′,N′-tetraacetic acid; FKBP-12, FK-506-binding protein 12; HEPES, N-(2-hydroxyethyl)pipera-zine-N′-(2-ethanesulphonic acid); HVA, high-voltage activated; HPLC, high-performance liquid chromatography; NMDAR, N-methyl-d-aspartate receptor; PCR, polymerase chain reaction; RT, reverse transcription; TEA, tetraethylammonium; VSCC., voltage-sensitive Ca2+ channel
 
In recent years it has become clear that Ca2+/calmodulin-dependent protein phosphatase 2B, or calcineurin is an important modulator of a wide range of critical functions in neurons. Calcineurin plays a major role in neuronal gene regulation, and appears to exert considerable control over genes involved with Ca2+ signaling and homeostasis (Carafoli et al., 1999; Genazzani et al., 1999; Graef et al., 1999). Moreover, increased activation or overexpression of calcineurin impairs memory (Mansuy et al., 1998), accelerates the decay of long-term potentiation (Winder et al., 1998), and increases vulnerability to apoptosis (Asai et al., 1999; Wang et al., 1999).

In turn, calcineurin may exert negative feedback over its own activation by inactivating specific Ca2+ sources, including voltage-sensitive Ca2+ channels (VSCCs) and N-methyl-d-aspartate (NMDA) receptors (NMDARs). However, this regulation is somewhat controversial. Although a number of studies suggest that calcineurin is important in Ca2+-dependent inactivation of VSCCs and NMDARs (Armstrong, 1989; Tong et al., 1995; Schuhmann et al., 1997; Lukyanetz et al., 1998) other reports have not reached a similar conclusion (Legendre et al., 1993; Krupp et al., 1996; Branchaw et al., 1997; Victor et al., 1997; Zeilhofer et al., 2000). One possible basis for conflicting results, at least for VSCCs, could be the differential expression of multiple VSCC types, or their association with disparate regulatory proteins in different cell types. There are at least five functional types of high-voltage-activated (HVA) VSCCs (L-, N-, P-, Q-, and R-type) in neurons (Fox et al., 1987; Bean, 1989; Fisher et al., 1990; Llinas et al., 1992; Mintz and Bean, 1993; Tsien et al., 1995) and it is well-documented that the mechanisms that regulate different VSCC types are highly selective, and moreover, can vary across cell types (Holz et al., 1986; Hescheler et al., 1987; Bean, 1989; Cox and Dunlap, 1992; Hille, 1994; Dolphin, 1995; Surmeier et al., 1995; Currie and Fox, 1997; Zamponi and Snutch, 1998; Ikeda and Dunlap, 1999). Even for a particular VSCC type, regulatory mechanisms may differ depending on state or environment. Regulation of the L-type VSCC by protein phosphorylation, for example, can depend on whether the full or truncated form of the α1C (Cav1.2) subunit is present (Hell et al., 1993a,b, 1996), or on whether L-type VSCCs are examined in native vs. non-native cell types (Zong et al., 1995).

It is less well-recognized, however, that calcineurin may also positively regulate some ion channels, including the GABAA receptor (Jones and Westbrook, 1997) and a renal Cl channel (Marunaka et al., 1998). Moreover, the possibility that calcineurin positively regulates VSCCs in some cell types has been raised recently. A series of preliminary studies in our laboratory revealed that the calcineurin inhibitor, FK-506, reduced rather than enhanced VSCC currents in cultured hippocampal neurons (Norris et al., 1999). A similar action of FK-506 on VSCCs also has been reported for arterial smooth muscle cells (Yasutsune et al., 1999). In addition, overexpression of calcineurin in transgenic mice enhances Ca2+ current density in cardiac myocytes (Yatani et al., 2001). Positive regulation of VSCCs by calcineurin in the hippocampus may be of particular interest, as calcineurin and multiple VSCC types are highly expressed in this brain structure (Fisher et al., 1990; Polli et al., 1991; Kuno et al., 1992; Llinas et al., 1992; Westenbroek et al., 1992; Hell et al., 1993a,b) and appear to be coupled functionally (Graef et al., 1999).

Here, we tested the possible positive modulation of VSCCs by calcineurin in hippocampal neurons using three distinct calcineurin inhibitors [the immunosuppressants FK-506 and cyclosporin A (CsA), and a calcineurin autoinhibitory peptide (CN-AIP)] and two negative controls (rapamycin and okadaic acid). Moreover, the effects of calcineurin inhibition on isolated VSCC subtype currents were examined to determine if hippocampal VSCC regulation by calcineurin is VSCC-type selective. Because L-type VSCC density and N-type VSCC regulation change with age in-culture (Porter et al., 1997; Blalock et al., 1999), somewhat analogously to changes with aging in vivo (Thibault and Landfield, 1996), we further examined the effect of age in culture on VSCC regulation by calcineurin.

EXPERIMENTAL PROCEDURES

Cell cultures
Primary hippocampal cell cultures, plated on 35 mm plastic culture dishes, were prepared from fetal pup tissue (embryonic day 18) obtained from pregnant Sprague–Dawley rats using slight modifications of the Banker and Cowan (1977) method, as previously described (Porter et al., 1997; Blalock et al., 1999; Brewer et al., 2001). Results for each condition are based on multiple dishes from at least two cultures. Generally, each rat pup provided enough tissue for two culture dishes. Most experiments were conducted on cells that were between seven and 10 days in vitro (DIV) in age (however, a few experiments were conducted on cells that were 13–14 DIV, see Results). For aging studies, sister cultures aged 14–17 DIV (2-week-old) and 28–31 DIV (4-week-old) were compared.

Electrophysiology

Glass electrodes Recording pipets consisted of glass capillary tubes (Drummond Scientific, Broomall, PA, USA) pulled on a horizontal micropipet puller (model P-87; Sutter Instruments, Novato, CA, USA). Whole-cell pipets were coated with polystyrene Q-dope and had a mean tip resistance of 2.16 ± 0.02 MΩ. Cell-attached patch pipets were coated with Sylgard (Dow Corning, Midland, MI, USA) and had a mean tip resistance of 2.7 ± 0.07 MΩ. All recording pipets were fire-polished immediately before recording (Corey and Stevens, 1983).

Recording solutions For whole-cell recordings of isolated HVA VSCC currents, external solution contained (in mM): 111 NaCl, 5 BaCl2, 5 CsCl, 2 MgCl2, 10 glucose, 10 HEPES, 20 tetraethylammonium (TEA) Cl, 0.01 6-cyano-7-nitroquinoxaline-2,3-dione disodium (CNQX) and 0.001 tetrodotoxin (TTX). pH was adjusted to 7.35 using NaOH and osmolarity adjusted to 330 mOsm using sucrose. Pipette solution for whole-cell recordings contained (in mM): 145 methane sulfonic acid, 10 HEPES, 3 MgCl2, 11 EGTA, 1 CaCl2, 5 MgATP, 13 TEA Cl, 0.1 leupeptin. pH was adjusted to 7.35 using CsOH and osmolarity adjusted to 320 mOsm using high-performance liquid chromatography (HPLC) grade H2O. This ratio of EGTA to Ca2+ buffers the intracellular Ca2+ concentration ([Ca2+]i) at levels below resting values (e.g., at <100 nM, Bers et al., 1994). To examine whole-cell Ca2+, rather than Ba2+, currents some experiments exchanged external BaCl2 for an equimolar amount of CaCl2. In one subset of these experiments, the internal solution was unchanged, while in other experiments, EGTA and CaCl were omitted and MgATP was replaced with an equimolar amount of 2NaATP.

For cell-attached patch recordings of multichannel activity, the external bath solution contained (in mM): 140 K+ gluconate, 3 MgCl2, 10 D-glucose, 10 EGTA, 10 HEPES. This solution, commonly used in single-Ca2+ channel studies, zeros the membrane and thus provides a convenient reference for setting the patch membrane potential (Fox et al., 1987; Fisher et al., 1990). pH was adjusted to 7.35 using KOH and osmolarity adjusted to 300 mOsm with HPLC grade H2O. The pipet solution consisted of (in mM): 20 BaCl2, 90 choline Cl, 10 TEA Cl, 10 HEPES. pH was adjusted to 7.35 using TEA-OH and osmolarity adjusted to 290 mOsm using sucrose. Prior to recording, the culture medium in each 35-mm dish was exchanged for 1.5 ml of external recording solution.

Data acquisition Recordings were obtained according to standard patch-clamp methods (Hamill et al., 1981) using an Axopatch 200A integrating patch-clamp amplifier (Axon Instruments, Foster City, CA, USA). Data were filtered at 2 kHz and digitized at 5 kHz. Voltage commands and data acquisition were controlled by pCLAMP software (Clampex, versions 6 and 7; Axon Instruments). All experiments were carried out at room temperature.

Whole-cell studies Prior to recording, junction potentials were nulled in the bath using the pipet offset control on the Axopatch 200A. Pipette capacitance was compensated. To estimate whole-cell membrane capacitance and pipet access resistance, a membrane current (filtered at 10 kHz, digitized at 91 kHz) was elicited at the beginning of each experiment with a 15-ms, 5-mV hyperpolarizing step from the holding potential (−70 mV). Current elicited by a 5-mV depolarizing pulse was equal in magnitude, but opposite in polarity. Due to their highly elaborate dendritic arbors, hippocampal neurons are not isopotential and exhibit capacitive current decay kinetics that are probably best fit by multiple exponential functions (Brown and Johnston, 1983; Johnston and Brown, 1983; Spruston et al., 1994). Therefore, we calculated whole-cell capacitance for each cell by integrating the area under the curve of the capacitive transient. The instantaneous peak current measured during the onset of the capacity transient was used to derive the pipet access resistance, which averaged 8.6 ± 0.14 MΩ, and was not significantly affected by the age of the cells or by the pharmacological agents used. Neurons in which the access resistance and/or the holding current changed dramatically during the course of an experiment were excluded from statistical analyses. Series resistance compensation, using the amplifier’s whole-cell correction parameters, was not performed because we, and others, have consistently found no significant differences in the shape or amplitude of activated currents before and after correction (Randall and Tsien, 1995; Porter et al., 1997; Blalock et al., 1999). For most experiments, the uncompensated error in the membrane potential during peak activation (i.e. voltage step to +10 mV) was estimated at 5–7 mV. Holding potential in all experiments was −70 mV. To generate current–voltage (IV) relationships, cells were stepped in 10-mV increments (10-s inter-step interval) from their holding potential at −70 mV, to +60 mV. Drug effects were assessed using a voltage protocol that stepped the membrane potential (for 150 ms) from its holding potential to the voltage that elicited essentially maximal/HVA VSCC current (either +10 mV when EGTA was included, or +20 mV when EGTA was excluded from the pipet solution). Unless otherwise stated, five successive current records (30-s interstep interval) were averaged to estimate maximal current for each cell. Online leak subtraction was carried out using a fractional (P/-5) method (as described in Porter et al., 1997 and Blalock et al., 1999), which consisted of five fractionally scaled hyperpolarizing subpulses. Current density (pA/pF) was calculated by dividing the mean whole-cell current (during a voltage step) by the membrane capacitance.

Under these recording conditions, whole-cell Ca2+ channel currents are typically followed by a long, post-repolarization Ca2+ current (e.g. Campbell et al., 1996; Porter et al., 1997). The source of this current remains controversial (cf. discussion in Thibault et al., 1993; Blalock et al., 1999), but the current appears too long to be due to space clamp artifact. Because it does not appear to alter the activated current and its source is not clear, it was not analyzed in the present studies.

Data obtained from IV curves were normalized and fit to a Boltzmann equation of the form:

equation M1

where I is the maximal current from the IV function, V is the voltage, V½ is the voltage of half-maximal activation, and k represents the steepness of the sigmoid curve. To measure inactivation of HVA Ca2+ currents, current amplitude sampled at 1 kHz during a 500-ms pulse was normalized to the peak current amplitude and fit with a double exponential function of the form:

equation M2

where 1/b = τfast, 1/d = τslow, A is the amplitude of each exponential component, and C is the steady-state asymptote. Fitted parameters were then compared across treatment groups using a z test.

Cell-attached multichannel patch studies Seal resistance in cell-attached patch studies averaged 21.9 ± 0.72 GΩ and was not altered by drug treatments. Patches with a seal resistance less than 15 GΩ were excluded from statistical analysis. For IV relationships, patches were stepped in 10-mV increments (10-s interstep interval) from their holding potential to +40 mV. Maximal current in each patch was achieved by stepping the patch membrane from its holding potential of −70 mV to +10 mV. A series of 15 steps to +10 mV was used to generate an average ensemble current for each patch. In prior studies with similar methods (e.g., Thibault et al., 1993; Thibault and Landfield, 1996; Porter et al., 1997; Brewer et al., 2001), we have observed that multichannel responses from the same cell are relatively consistent, and consequently, 15 step-ensembles generate results that are highly similar to those from 30- or 45-step-based ensembles. The relatively large patches employed here, which can contain numerous channels (Thibault and Landfield, 1996; Porter et al., 1997; Brewer et al., 2001), appear to confer stochastic stability. Each depolarization step was separated by 15 s to prevent channel inactivation from confounding the ensemble current amplitude. In each patch, currents generated from a series of hyperpolarizing pulses (equal in amplitude to the depolarizing steps) were averaged and added offline to ensemble currents to generate leak-subtracted waveforms (Fetchan version 6, Axon Instruments). Current density (pA/μm2) for each patch was derived by dividing the average ensemble current by the patch area. The patch area, which is inversely proportional to the pipet resistance, was estimated for each patch using the equation a = 12.6(1/R+0.018), where a is the patch area and R is the pipet resistance (Sakmann and Neher, 1983).

Drugs and drug delivery FK-506, CsA, rapamycin, and okadaic acid were dissolved in 100% ethanol. These drugs were then diluted ≥1000×, such that the ethanol concentration in the recording solutions did not exceed 0.1%. Higher concentrations of FK-506 and CsA (>20 μM) typically required sonication to go into solution. CN-AIP was dissolved in dH2O (at 1000×). All drugs were purchased from Calbiochem (La Jolla, CA, USA), except the FK-506-binding protein (FKBP-12), and okadaic acid, which were obtained from Sigma (St. Louis, MO, USA). For incubation studies, drugs were added to the cell culture medium 2–4 h prior to recording. Culture medium then was exchanged for drug-free recording medium immediately prior to the experiment. For intracellular application studies, drugs were dissolved in the pipet solution. In channel antagonist and aging studies, FK-506, dissolved in external recording medium, was delivered to the recording site by free diffusion from a low-resistance glass (weeper) pipet (Blalock et al., 1999). In these latter experiments, CNQX and TTX were included in the weeper pipet, but excluded from the external solution. Consequently, full diffusion from the weeper pipet could be verified by the absence of a fast Na+ current transient and the paucity of synaptic events in the whole-cell current record.

mRNA analysis

Neuron collection For mRNA studies, pyramidally shaped neurons from 2-week and 4-week-old sister cultures were aspirated into glass pipets (1–1.5 MΩ tip resistance), filled with 5 μl of standard whole-cell internal recording solution. For each culture dish, a single pipet was used to aspirate one neuron at a time until approximately 15 neurons were collected. With this method, the entire neuronal cell body and a portion of the neuron’s proximal neurites were cleanly extracted with minimal collection of extraneous surrounding glial or neuronal processes. Pyramidally shaped neurons of the type recorded were selected for collection. For statistical analyses of mRNA data, n equaled the number of culture dishes analyzed in each age group.

Following collection, neurons were immediately pipetted into a pre-chilled microcentrifuge tube containing RNase inhibitor (1 U/μl) and random hexamers (0.25 mM). This cellular lysate was heated for 10 min at 70°C, cooled on ice and followed by reverse transcription (RT) for cDNA synthesis. The RT was carried out in a 20-μl reaction containing 1× polymerase chain reaction (PCR)II buffer (Perkin Elmer, CA, USA), 3.75 mM MgCl2, 0.5 mM dNTPs and 50 U MMLV reverse transcriptase (Gibco BRL). The RT reaction mixture was incubated at 42°C for 1 h followed by heating at 95°C for 10 min and then stored at −20°C until further PCR analysis.

Quantitative real-time PCR Quantitation of relative mRNA content was performed using an ABI prism 7700 sequence detection system (Applied Biosystems, CA, USA) and the TaqMan methodology, which utilizes the 5′ nuclease activity of Taq DNA polymerase to generate a real-time quantitative DNA assay. In brief, mRNA-specific oligonucleotide probes (TaqMan probes) containing 5′-fluorescent reporter and 3′-quencher dyes were designed and used for the extension phase of the PCR. The degradation and release of the reporter dye (i.e. FAM) results in fluorescence at 518 nm, which is monitored during the complete amplification process.

All the cDNA samples from the 2-week-old (n = 13) and 4-week-old (n = 14) groups were analyzed simultaneously by real-time PCR, with each reaction run in duplicate. The PCR solution was based on the Single Reporter real-time PCR protocol, as described in the ABI 770 system’s manual. For each mRNA species, the amount of reverse transcriptase used for the PCR was optimized to ensure that the results fell within the standard curve. To normalize the amount of calcineurin and calmodulin cDNA in each RT sample, the ribosomal RNA 18S level was measured using a 100-fold dilution of the RT as the template for real-time PCR. The primers for 18S were: forward, 5′-AGTCCCTGCCCTTTGTACACA-3′; reverse, 5′-GATCCGAGGGCCTCACTAAAC-3′; and the TaqMan probe was: 6FAM-CGCCCGTCGCTACTACCGATTGG-TAMRA (nt 1684–1752, GenBank X01117). For the calcineurin message, 1 μl of RT was used as the template in a 50-μl PCR reaction mixture. The primers for calcineurin were: forward, 5′-TCCCGGTGACTGGAGATGTC-3′; reverse, 5′-TTTCAC-CACCCTGTCCGTAGT-3′; the TaqMan probe was: 6FAM-CCCAAGGCGATTGATCCCAAGTTGT-TAMRA (nt 193–264, GenBank D90036). For calmodulin, 0.5 μl of RT was used as the template in each PCR reaction and the primers were: forward, 5′-GTCTGTTCTGGTCTCGGAAACC-3′; reverse, 5′-TGAATTCTGCGATCTGCTCTTC-3′; the Taq-Man probe was: 6FAM-CCTTGCAGCATGGCTGACCAACTGA-TAMRA (nt 34–110, GenBank M17069).

PCR quantitative analysis
Real-time PCR quantitation was performed using the ABI 7700 system’s software. At the completion of PCR (a total of 40 cycles), the amount of target message in each reaction was determined from the detection threshold cycle number (CT), which is inversely correlated with the abundance of the message’s initial level. The CT was then converted to relative quantity by normalizing to a standard curve. A standard curve was constructed by running PCR simultaneously using 10-fold serial dilutions of a stock cDNA template containing known quantities of message, as described in the ABI 7700 system’s manual.

Statistical comparisons VSCC current records were analyzed quantitatively using Clampfit 6.0 software (Axon Instruments) and all statistical analyses were performed using Statview 5.01. In most studies, the significance of drug effects was determined by analysis of variance (ANOVA). For studies utilizing pre- and post-drug application measures, effects were determined by repeated measures ANOVA. Post hoc analyses were performed using Scheffe’s F-test.

RESULTS

Effects of calcineurin and protein phosphatase 1 and 2A inhibition on HVA VSCC currents
In the first series of studies, we asked the question of whether calcineurin activity was necessary for HVA VSCC function in hippocampal neurons, and if so, whether a calcineurin/protein phosphatase cascade (Cohen, 1988) mediated this action. FK-506 and CsA are structurally distinct immunosuppressive agents that specifically inhibit calcineurin activity by binding to separate, endogenously expressed immunophilins. FK-506 binds to FKBP-12, while CsA binds to cyclophilin A (Snyder et al., 1998). Okadaic acid, on the other hand, is a relatively specific inhibitor of protein phosphatases 1 and 2A and exhibits little potency toward calcineurin at drug concentrations of ≤1 μM (Cohen, 1991).

Hippocampal neurons were incubated for 2–4 h in medium containing either FK-506 (at three concentrations: 0.5, 5, or 50 μM), CsA (20 μM) or okadaic acid (1μM) (see Experimental procedures). These initial studies used drug incubation instead of acute application to ensure equilibrium with immunophilins, which may be limiting in brain (Kung and Halloran, 2000) and other binding sites. Incubation across multiple concentrations of FK-506, revealed a concentration-dependent reduction in whole-cell VSCC Ba2+ current density relative to the respective ethanol-incubated control cells (n = 11–15 control cells at each of the three FK-506 concentrations) [overall ANOVA: F(3,75) = 17.57, P < 0.001 (Fig. 1A, B)]. Post hoc analyses showed that 0.5 μM (n = 9) induced a small, non-significant reduction (90 ± 5% of control, P = 0.11), 5 μM (n = 16) induced an ~25% decrease (75 ± 5% of control, P < 0.05), and 50 μM (n = 13) resulted in ~50% inhibition (52 ± 4% of control, P < 0.001, n = 13). Higher concentrations of FK-506 appeared to induce no further inhibition, although they were difficult to study because of the drug’s lipophilicity. At none of the employed concentrations did FK-506 significantly alter the voltage dependence of VSCCs (Fig. 1C). Overall, the half-maximal activation voltage for FK-506 studies was −7 ± 0.76 mV.

Fig. 1Fig. 1
FK-506 and CsA, but not okadaic acid, inhibited whole-cell HVA Ba2+ current. (A, D) Representative examples of averaged VSCC traces (from five successive sweeps) recorded during repeated voltage steps from −70 mV to +10 mV in individual vehicle (more ...)

CsA at 20 μM also significantly reduced VSCC current density by approximately 25% (CsA, 10.4 ± 0.7 pA/pF, n = 13; control, 13.8 ± 1.2 pA/pF, n = 14) [F(1,25) = 5.372, P < 0.05], without altering the IV relationship (Fig. 1D–F). In contrast, cells treated with okadaic acid (1 μM) (n = 12), did not show a significant reduction in maximal current. However, they exhibited a shift to the left in the IV curve, relative to control cells (n = 12) (control V½ = −11.92 mV; okadaic acid V½ = −20.34 mV; z = 7.77, P < 0.05; Fig. 1G), indicating that okadaic acid increased voltage sensitivity. Thus, unlike FK-506 and cyclosporin, okadaic acid tends to enhance, rather than inhibit VSCC current. These results indicate that calcineurin activity, but not protein phosphatases 1 and 2A activity, appears to be required for a significant fraction of VSCC current in hippocampal neurons.

Specificity of effects on VSCC current for inhibition of calcineurin
Although both FK-506 and CsA are considered to be specific calcineurin inhibitors, both act as immunophilin ligands and also are immunosuppressants with non-calcineurin-mediated actions (Snyder et al., 1998). In addition, CsA stabilizes mitochondrial Ca2+ homeostasis (Connern and Halestrap, 1994; Ankarcrona et al., 1996; Friberg et al., 1998). Therefore, to confirm that the inhibition of VSCC current in Fig. 1 was specific to drug actions on calcineurin, we also examined the effects of a specific negative control, rapamycin, and a specific positive control, calcineurin-AIP on HVA Ba2+ currents. Rapamycin is an immunosuppressant that is similar in structure to FK-506 and competes for binding to FKBP-12 (Dumont et al., 1990). However, unlike the FK-506/FKBP-12 complex, the rapamycin/FKBP-12 complex does not bind to and inhibit calcineurin (Snyder et al., 1998). Thus, rapamycin is an advantageous agent for separating FK-506’s actions on immunophilins from its actions on calcineurin. CN-AIP is similar to the C-terminal region of the calmodulin-binding domain of calcineurin (Hashimoto et al., 1990; Perrino et al., 1995) and acts as a highly specific competitive inhibitor of calcineurin A catalytic activity (Sagoo et al., 1996).

In this series of studies, agents were applied intracellularly to ensure equal cellular access of the compounds being compared. Hippocampal neurons were dialyzed, via the whole-cell recording pipet, with either 5 μM FKBP-12 alone (control, n = 29), or FKBP-12 in combination with either FK-506 (n = 25), rapamycin (n = 17), or FK-506 together with rapamycin (n = 7). Thus, FKBP-12 was present in all experimental conditions. All immunophilin ligand concentrations were set at 5 μM. Recordings were begun after three stable current records were obtained, at approximately 10 min after establishing the whole-cell configuration. At this point, five successive current traces generated by a 150-ms step from −70 mV to +10 mV (interpulse interval 30 s) were averaged. Cells treated with FK-506/FKBP-12 exhibited a significant reduction in whole-cell Ba2+ current density, relative to all other conditions [F(3,73) = 8.679, P < 0.001] (Fig. 2A, B). Rapamycin/FKBP-12 differed only from the FK-506/FKBP-12 group (P < 0.01) and had no inhibitory action on VSCC current. Moreover, when rapamycin was included in the pipet with FK-506/FKBP-12, the inhibitory effects of FK-506 were blocked (post hoc comparison between the FK-506/FKBP-12 and the FK-506/FKBP-12+ rapamycin groups, P < 0.05). These findings suggest that rapamycin at 5 μM was able to bind FKBP-12, and compete with FK-506. They also indicate that ligand binding to FKBP-12 alone is not sufficient for VSCC inhibition. In contrast to FK-506/FKBP-12, rapamycin/FKBP-12 induced a small leftward shift in the IV function, such that half-maximal activation occurred at −15.79 ± 0.81 mV as opposed to −10.63 ± 0.7 mV for FKBP-12 alone cells (z = 4.82, P < 0.05) and −11.52 ± 1.21 for FK-506/FKBP-12 cells (z = 2.92, P < 0.05) (Fig. 2C). Thus, rapamycin and FK-506, which differ primarily in their abilities to inhibit calcineurin, also have strikingly different effects on VSCC current. Rapamycin tended to enhance rather than inhibit VSCC current. Although it inhibits calcineurin activity through a different mechanism from FK-506, CN-AIP (100 μM) also significantly reduced whole-cell Ba2+ current density [F(1,20) = 8.627, P < 0.01] (Fig. 2D, E). As with FK-506 and CsA, CN-AIP did not substantially alter the voltage dependence of VSCCs (control V½ = −11.73 ± 0.55 mV, CN-AIP V½ = −13.66 ± 1.5) (Fig. 2F). Together, the results indicate that inhibition of calcineurin is the necessary and sufficient factor in the inhibition of HVA VSCC current by these multiple agents.

Fig. 2Fig. 2
Reduction of HVA Ba2+ current was specific to calcineurin inhibition. Whole-cell currents were recorded in neurons dialyzed with either FKBP-12 alone (5 μM), FKBP-12/FK-506 (Rap; 5 μM), FKBP-12/rapamycin (5 μM), or FKBP-12/FK-506 (more ...)

FK-506 inhibits maximal VSCC current amplitude, but does not affect Ca2+-dependent inactivation
In the previous experiments, Ba2+ rather than Ca2+ was used as the charge carrier, and the internal [Ca2+] was buffered to the low nanomolar range by the Ca2+ chelator EGTA. As such, these conditions may underestimate or confound the role of a Ca2+/calmodulin-dependent enzyme, such as calcineurin, in VSCC regulation. However, although maximal catalytic activity of calcineurin requires a substantial elevation in cytosolic Ca2+ (e.g. to the micromolar range), calcineurin can still exhibit significant phosphatase activity in the presence of low Ca2+ and/or EGTA (Stewart et al., 1982; Liu et al., 1991; Stemmer and Klee, 1994). Nonetheless, it seems important to determine whether calcineurin regulation of VSCCs is dependent on the charge carrier used, or on whether internal Ca2+ is buffered at a low level. In the following experiments, therefore, external BaCl was replaced with an equimolar amount of CaCl. In Fig. 3A, the pipet solution still contained 11 mM EGTA, while the experiments in Fig. 3B, C were conducted with no EGTA in the pipet. As described above for Fig. 2, cells were dialyzed with either FKBP-12 alone (5 μM) or FKBP-12 along with FK-506 (both at 5 μM) and maximal VSCC current was measured. As shown in Fig. 3A, replacing external Ba2+ with Ca2+ did not appear to impair FK-506’s capacity to inhibit VSCC current. Similar to the results in Fig. 2, maximal current amplitude recorded from cells dialyzed with FK-506/FKBP-12 (n = 10) was significantly reduced compared to those dialyzed with FKBP-12 alone (n = 12) [7.75 ± 0.64 vs. 10.33 ± 0.78 pA/pF, F(1,20) = 6.2, P < 0.05].
Fig. 3Fig. 3
FK-506 reduced HVA Ca2+ currents, but did not alter VSCC inactivation. Extracellular Ba2+ was replaced with equimolar Ca2+ and neurons were dialyzed with either FKBP-12 alone (5 μM) or FKBP-12/FK506 (5 μM) as described in Fig. 2. FK-506 (more ...)

Fig. 3B shows that removal of EGTA from the pipet solution also did not prevent FK-506 from inhibiting Ca2+ current through VSCCs [FK-506/FKBP-12 = 13.35 ± 1 pA/pF, n = 12; FKBP-12 alone = 16.97 ± 1.27 pA/pF, n = 14; F(1,24) = 4.8, P < 0.05]. The exclusion of EGTA from the pipet solution also gave rise to VSCC currents that exhibited substantial inactivation. That is, with EGTA present in the pipet solution (Fig. 3A), Ca2+ current dropped by only 8 ± 2% from its peak amplitude to its minimum amplitude at the end of the 150-ms voltage step. However, cells recorded without EGTA in the pipet showed a much greater reduction (34 ± 3%) in VSCC current during the voltage step [F(1,21) = 63.9, P < 0.001], indicating considerable Ca2+-dependent inactivation in EGTA-free cells.

To examine whether FK-506 altered Ca2+-dependent inactivation, a 500-ms depolarizing pulse to +20 mV was applied to cells in the absence of internal EGTA (Fig. 3C). The time course of VSCC inactivation could be well-fit using a double exponential function (see Experimental procedures). The results showed that, in contrast to its effects on maximal VSCC amplitude, FK-506 did not significantly alter the rate of VSCC inactivation (Fig. 3C, right panel). However, both the slow and the fast time constants were slightly reduced in the FK-506/FKBP-12 group. Although inactivation kinetics in dendritically ramified neurons can be confounded by multiple compartments (Brown and Johnston, 1983; Johnston and Brown, 1983; Spruston et al., 1994), the absense of an effect on either exponential function indicates that, in hippocampal neurons, FK-506 can reduce whole-cell Ca2+, as well as Ba2+ currents, without appreciably altering the inactivation kinetics of HVA VSCCs. Interestingly, cells without EGTA exhibited larger Ca2+ currents relative to cells in which Ca2+ was buffered [F(1,24) = 18.42, P < 0.001] (Fig. 3B vs. Fig. 3A). Although the fraction of current inhibited by FK-506 in these two conditions was similar, the results are at least consistent with the possibility that elevating intracellular Ca2+ can, under some circumstances, enhance Ca2+ channel activity, perhaps via a calcineurin mechanism.

FK-506 inhibits VSCC activity in multichannel cell-attached patches
As noted, the whole-cell recording configuration introduces problems that may confound measures of VSCC activity. For example, hippocampal neurons exhibit extensive dendritic arbors and may be subject to space clamping errors (Brown and Johnston, 1983; Johnston and Brown, 1983; Spruston et al., 1994). In addition, cytosolic components, including calcineurin, immunophilins, kinases, or other phosphatases critical for regulating VSCCs may be lost when cells are dialyzed with the whole-cell pipet recording solution. Analysis of single-channel-level activity in the cell-attached patch-clamp configuration circumvents space clamp problems and avoids membrane rupture, thereby preserving the integrity of cytosolic constituents. Therefore, to test whether FK-506-mediated VSCC inhibition was influenced by the whole-cell recording configuration, VSCC activity also was examined in cell-attached patches in hippocampal neurons after FK-506 treatment.

To enhance access of FK-506 to channels in the patch, cells were incubated in FK-506 (50 μM) for at least 2 h. Neurons exposed to FK-506 (n = 12) exhibited a significant decrease in cell-attached VSCC activity (~45%) relative to control patches (n = 8) [F(1,18) = 9.64, P < 0.01] (Fig. 4A, B), with no change in the voltage dependence of VSCCs (data not shown). Thus, inhibition of VSCCs by FK-506 was not limited to the whole-cell recording configuration. Moreover, the magnitude of inhibition was relatively similar in both whole-cell (Fig. 1) and cell-attached patch recording modes (Fig. 4).

Fig. 4Fig. 4
FK-506 reduced HVA Ba2+ current activity in multichan-nel cell-attached patches. (A) Representative waveforms of five current traces, and the average ensemble current (lowest trace) generated from 15-step depolarizations in multichannel cell-attached (more ...)

Isolation of specific VSCC types
Because saturating concentrations of calcineurin inhibitors did not block all HVA VSCC current, we tested whether calcineurin’s apparent enabling or enhancing action might be selective for only some VSCC types. HVA VSCCs in neurons comprise at least five identified functional types, L-, N-, P-, Q- and R-type (Fox et al., 1987; Bean, 1989; Fisher et al., 1990; Llinas et al., 1992; Mintz and Bean, 1993; Tsien et al., 1995) generally corresponding to different α1 subunits encoded by specific genes (Birnbaumer et al., 1994; Catterall, 1995). The L-type VSCCs are blocked highly selectively by dihydro-pyridines, such as nimodipine (Fox et al., 1987; Fisher et al., 1990), whereas N-type and P/Q-type channels are fully and selectively blocked by snail toxins (GVIA and MVIIC conotoxins, respectively) (Mintz and Bean, 1993; Randall and Tsien, 1995; McDonough et al., 1996). Using these specific channel antagonists to isolate VSCC current components, we tested whether any antagonists occlude the effects of FK-506. In these experiments, whole-cell maximal Ba2+ current records were obtained every 30 s and specific VSCC antagonist drugs and FK-506 were applied locally to neurons via diffusion from a ‘weeper’ pipet (see Experimental procedures). Drug application was begun after at least three stable current traces were recorded, which was usually between 5 and 10 min after establishing the whole-cell configuration. Current records for the analysis of drug effects were obtained at the end of a drug application, when current amplitudes were stable. Ten minutes of ethanol vehicle alone (n = 9) induced no inhibition or rundown of VSCC current (Fig. 5A, open circles). FK-506 (50 μM) was delivered after 3–5 min application of either 0.1% ethanol alone (Fig. 5A, filled circles), nimodipine (10 μM) (Fig. 5B), or a solution containing both ω-conotoxins, MVIIC (3 μM) and GVIA (1 μM) (Fig. 5C). The occurrence and/or the extent of VSCC inhibition by FK-506 was highly dependent on the components of VSCC current available during FK-506 application [F(3,31) = 36.7, P < 0.001]. When delivered after 0.1% ethanol, FK-506 induced a slowly developing (usually within 2–3 min) inhibition of VSCC activity, that reached maximal, asymptotic levels by 6–7 min (current remaining after FK-506:62 ± 3% of prior ethanol baseline, n = 16, P < 0.001) (Fig. 5A, filled circles). This FK-506 effect was not reversible upon wash for up to 10 min (data not shown). In contrast, prior inhibition of the L-type fraction of VSCC current with a saturating concentration of nimodipine (n = 6) completely occluded further inhibition by FK-506 (current remaining after FK-506:94 ± 6% of prior nimodipine baseline, P > 0.05) (Fig. 5B). Conversely, when VSCC current was first inhibited by the N- and P/Q-type VSCC blockers, MVIIC and GVIA (n = 4), such that the remaining HVA current comprised L-type and R-type components (Randall and Tsien, 1995; McDonough et al., 1996), FK-506 induced a significantly larger fractional reduction in the remaining L-type-enriched current (current remaining after FK-506:35 ± 2% of prior conotoxin baseline, P < 0.01, Fig. 5C). Thus, L-type VSCC activity was required for the inhibitory effect of FK-506 on VSCC current (Fig. 5B), and, in L-type channel-enriched HVA VSCC current, FK-506 blocked proportionally more VSCC current (Fig. 5C). Consequently, the L-type VSCC appears to be both necessary and sufficient for the enabling/enhancing action of calcineurin on Ca2+ current.
Fig. 5Fig. 5
FK-506 selectively inhibited L-type VSCC current. (A–C) Left panels. Time course plots of mean (± S.E.M.), normalized VSCC currents of neurons exposed to acute external perfusion of 0.1% ethanol alone (A, open circles), 50 μM FK-506 (more ...)

Age-dependent effects of FK-506 on VSCC current
Brain aging has been associated with an increase in the density of available L-type VSCCs and L-VSCC current in hippocampal neurons (Landfield et al., 1989; Disterhoft et al., 1993; Thibault and Landfield, 1996). In an apparently analogous manner, L-VSCC currents and L-type channel density undergo an age-dependent increase in cultured hippocampal neurons, albeit over a much shorter time course. A major increase in L-VSCC current occurs between 1 and 4 weeks in culture (Porter et al., 1997; Blalock et al., 1999). During this same time frame, the inhibitory regulation of N-type and P/Q type VSCCs by the membrane-delimited G-protein-coupled pathway (Holz et al., 1986; Hescheler et al., 1987; Bean, 1989; Cox and Dunlap, 1992; Hille, 1994; Dolphin, 1995; Currie and Fox, 1997; Ikeda and Dunlap, 1999) decreases substantially (Blalock et al., 1999). Therefore, it seemed of considerable interest to determine whether the selective regulation of L-VSCCs by calcineurin also changes with weeks in culture. To answer this question, neurons from 2-week- and 4-week-old sister cultures were perfused with FK-506, under conditions similar to those outlined for studies in Fig. 5. However, rather than using a saturating 50-μM concentration of FK-506, which appears to block all L-type VSCC activity (Fig. 5A, B), and consequently might obscure differences in the sensitivity or content of calcineurin, cells in the present experiment were perfused with 5 μM FK-506. Although a significant overall reduction in whole-cell Ba2+ current was found across both age groups after FK-506 application [F(1,14) = 28.6, P < 0.001], the extent of current modulation (measured as percent of baseline) was significantly greater (P < 0.05) in 4-week-old (current remaining after FK-506:73 ± 7% n = 8), relative to 2-week-old cells (current remaining after FK-506:90 ± 7%, n = 8) (Fig. 6A, B). In neither age condition did whole-cell current amplitude change by the end of a 10-min application of 0.1% ethanol vehicle alone (2 weeks, 99 ± 4%, n = 8; 4 weeks, 102 ± 7%, n = 6, data not shown). Extracellular application of 5 μM FK-506 did not reduce HVA current as much in younger cells in the present experiment as in the study shown in Fig. 1. This may be related to the extended incubation used in the Fig. 1 study or to normal variability among different cultures. Nevertheless, the present experiment compared sister cultures and the results reveal a greater FK-506 effect as a function of age in culture. Thus, as L-type VSCC density increases, the calcineurin-dependent component of total HVA VSCC current appears to increase concomitantly. However, the possibility of enhanced sensitivity of calcineurin to FK-506 cannot be ruled out by the present data.
Fig. 6Fig. 6
Effects of age in culture on VSCC inhibition by FK-506 and calcineurin mRNA level. (A) Representative time plots of normalized VSCC currents from a representative 2-week (open circles) and a representative 4-week-(filled circles) old neuron during acute (more ...)

Relative expression of calcineurin mRNA is elevated in older neurons
Hippocampal aging, across months in vivo (Herman et al., 1998; Chen et al., 2000) or weeks in culture (Porter et al., 1997), also is associated with an increase in mRNA for the Cav1.3 (α1D) pore-forming subunit of the L-type VSCC. Because one explanation for the greater effect of 5 μM FK-506 on VSCC current could be an up-regulation of calcineurin concomitantly with the L-type VSCC increase, we tested whether mRNA levels for the catalytic subunit of calcineurin (calcineurin A) also increase with age in culture. Neurons were collected individually from culture dishes (15 neurons/dish) of 2-week- and 4-week-old cells. mRNA for calcineurin was amplified and quantified using real-time PCR (see Experimental procedures). Semiquantitative estimates (normalized to 18S mRNA) of calcineurin mRNA levels in the 4-week-group (n = 14) were more than 2-fold greater than those of their 2-week-old counterparts (n = 13) [F(1,25) = 5.36, P < 0.05] (Fig. 6C). However, no significant differences in normalized mRNA content for calmodulin were observed (data not shown), indicating that age-related changes in calcineurin expression may be relatively selective.

DISCUSSION

Channel type-specific enhancing action of calcineurin in hippocampus
As discussed, a number of previous studies have found evidence that calcineurin mediates aspects of negative feedback regulation of VSCCs (Armstrong, 1989; Schuhmann et al., 1997; Lukyanetz et al., 1998; Burley and Sihra, 2000), in particular, of N-type VSCCs (Zhu and Yakel, 1997; Lukyanetz et al., 1998). Moreover, in some cells, D2 dopamine receptors apparently inhibit L-type VSCCs via a calcineurin-mediated pathway (Hernandez-Lopez et al., 2000). Conversely, the present studies provide the first evidence, to our knowledge, of a positive regulatory mechanism of calcineurin on VSCCs in neurons. No indication of an action of calcineurin on Ca2+-dependent inactivation was seen in this preparation (Fig. 3). Thus, this finding appears to extend the class of channel type-specific VSCC regulatory mechanisms (e.g., Bean, 1989; Hille, 1994; Dolphin, 1995; Surmeier et al., 1995; Zamponi and Snutch, 1998; Ikeda and Dunlap, 1999) to include an enhancing/enabling action of calcineurin on L-type VSCCs.

In the present work, a series of disparately acting calcineurin inhibitors, phosphatase inhibitors, and immunophilin ligands were compared to show that the common feature among agents that suppressed VSCC current was calcineurin inhibition. That is, these studies found strong evidence that calcineurin inhibition is both necessary (e.g. the similar FKBP-12 immunophilin ligand, rapamycin, which does not inhibit calcineurin, did not inhibit VSCC current) and sufficient (CN-AIP, which inhibits calcineurin but does not bind to immunophilins, also inhibited VSCC current) for mediating inhibition of VSCC current (Fig. 2). Consistent with these results, the PP1/PP2A inhibitor, okadaic acid, did not suppress VSCC current (Fig. 1). Finally, the effect of calcineurin inhibition did not depend on the conditions of the whole-cell recording configuration, as VSCC inhibition was of relatively similar magnitude in the cell-attached patch configuration (Fig. 4). Together, the results indicate clearly that inhibition of VSCCs by these agents specifically required calcineurin inhibition.

The enhancing effect of calcineurin also was targeted selectively to one type of VSCC, the L-type. The functional availability of L-type VSCCs appeared to be both necessary (i.e., nimodipine, the specific L-type antagonist, blocked all action of FK-506 on VSCCs) and sufficient (in the absence of N-, and P/Q-type VSCC activity, the relative FK-506 effect on L-type current was considerably greater) for the full manifestation of FK-506-mediated inhibition of VSCC currents in hippocampal neurons. Thus, the L-type VSCC was the only VSCC that was dependent on some degree of calcineurin phosphatase activity. Moreover, calcineurin-inhibition did not shift VSCC voltage dependence but instead appeared to block fully the activation of L-VSCCs. These data imply that calcineurin action is essential for enabling L-type VSCC availability in hippocampal neurons. However, it is also possible that, in addition to enabling availability, calcineurin enhances L-type VSCC activity through other processes (e.g. increased open probability).

Magnitude of the L-type current
Previous studies have found that L-type currents account for ~20–50% of somal VSCC currents in brain neurons (Fisher et al., 1990; Mintz and Bean, 1993; Randall and Tsien, 1995; Blalock et al., 1999) (although the conductance of L-type VSCCs, in particular, may vary depending on whether Ca2+ or Ba2+ is the charge carrier (Church and Stanley, 1996). The L-type component of HVA Ba2+ current increases with age in culture from ~30% in younger (1–2-week-old) to nearly 50% of current in older (3–4-week-old) hippocampal cultured neurons (Porter et al., 1997; Blalock et al., 1999). However, there is significant variability across cultures, such that comparisons across ages are generally carried out in sister cultures (see Experimental procedures). In the present study, a slightly larger L-type component appeared to be present in younger cells than in earlier studies. This is indicated by the >40% of current blocked by nimodipine in younger neurons (Fig. 5B), which was similar to the ~38% blocked by 50 μM FK-506 under similar application conditions (Fig. 5A). Moreover, in one experiment of the present studies, 50 μM FK-506 suppressed ~50% of VSCC current in younger neurons (Fig. 1). Conceivably, the prolonged incubation with FK-506 in the Fig. 1 study could have resulted in inhibition of a small residual amount of L-type current not readily blocked by acute application of nimodipine or FK-506. Alternatively, the differences between these studies could be due to normal variability across cultures.

Calcineurin sensitivity to inhibitors and [Ca2+]i
Studies with calcineurin inhibitors in neuronal tissues (e.g., Tong et al., 1995; Lukyanetz et al., 1998; Zeilhofer et al., 2000) often have employed drug concentrations that are high relative to the IC50 found in in vitro biochemical studies (Liu et al., 1991) or peripheral cell assays (Kung and Halloran, 2000). The relatively high effective concentrations of FK-506 in the present study (Fig. 1) also exceeds those seen in in vitro and peripheral assays. This discrepancy may relate to the recent observation that, contrary to the general assumptions, immunophilins in many tissues, particularly brain, are limiting for calcineurin inhibition by FK-506 and CsA (Kung and Halloran, 2000). This limitation might somehow result in an altered IC50 (e.g., immunophilins may be less accessible in neurons). Alternatively, there may be endogenous competitors or multiple compartments of calcineurin, with varying accessibility.

Maximal catalytic activity of calcineurin requires a large rise in [Ca2+]i (to the μM range) and subsequent binding to Ca2+/calmodulin (Stemmer and Klee, 1994). However, many of the present studies were performed in neurons in which Ba2+ was the charge carrier and/or [Ca2+]i was buffered by EGTA to near-resting (< 100 nM) concentrations (Bers et al., 1994). Thus, [Ca2+]i presumably did not rise substantially during current influx. Nevertheless, calcineurin also shows some phosphatase activity at resting or low [Ca2+]i levels, stimulated by Ca2+ bound to the calcineurin B subunit of the calcineurin holoenzyme. This activity is seen in the absence of calmodulin or in the presence of EGTA (Stewart et al., 1982; Liu et al., 1991; Stemmer and Klee, 1994). Therefore, it can be concluded that relatively low levels of calcineurin activity at near-resting [Ca2+]i can mediate the enabling action of calcineurin on L-type VSCCs. However, EGTA is a relatively slow Ca2+ buffer and the possibility that intracellular Ca2+ release and/or capacitive Ca2+ entry increased calcineurin activity above anticipated levels cannot be ruled out. Finally, the preliminary evidence that Ca2+ currents were larger when [Ca2+]i was not buffered (Fig. 3A,B) raises the intriguing possibility that, while low or basal levels of [Ca2+]i may be able to maintain the basic enabling effect of of calcineurin on L-type VSCCs, elevating [Ca2+]i, and in turn, increasing calcineurin activity, may recruit previously unavailable L-type VSCCs or enhance channel open probability. Clearly additional studies will be required to clarify the Ca2+ dependence of calcineurin’s actions on L-type VSCCs in hippocampal neurons.

Channel enablement by dephosphorylation
Higher phosphorylation/dephosphorylation ratios appear to positively modulate ion channel and ionotropic receptor functions under many conditions (Armstrong, 1989; Artalejo et al., 1992; Catterall, 1995; Dolphin, 1996). However, phosphorylation at some sites also can decrease ion channel activity. For example, protein kinase G-dependent phosphorylation of the L-type α1C subunit at position Ser (533) inhibits L-VSCC activity (Jiang et al., 2000). Moreover, calcineurin has been found to enhance the activity of other ion channel types (Jones and Westbrook, 1997; Marunaka et al., 1998). Conceivably, therefore, calcineurin activity might enable L-type VSCCs by dephosphorylating channel sites at which phosphorylation during channel opening contributes to inactivation. This could imply a rapid and continuous modulation of L-type VSCC availability that might require a close physical juxtaposition and/or complexing of calcineurin and L-VSCCs. Interestingly, there is some evidence that calcineurin is localized on neuronal membranes in close association with VSCCs (Lukyanetz, 1997). Moreover, there is increasing evidence of physical and functional coupling of protein phosphatases with intracellular Ca2+ channels (Cameron et al., 1995) or L-type Ca2+ channels (Davare et al., 2000). However, the possibility that calcineurin acts on L-type VSCCs indirectly through some other enzyme pathway obviously cannot be ruled out.

Age-dependent changes and functional implications
The present studies found that VSCC current was more sensitive to inhibition by 5 μM FK-506 in older than in younger cultured neurons (Fig. 6A, B). This might reflect an increase in calcineurin content, potentially as a required coordinated up-regulation with an increasing availability of L-type channels (Porter et al., 1997; Blalock et al., 1999). The relative age-related increase in calcineurin A mRNA content (Fig. 6C) is at least preliminarily consistent with this possibility. On the other hand, the results could also be accounted for by a heightened calcineurin sensitivity to FK-506 and/or an age-dependent increase in immunophilin abundance, among other possibilities.

Interestingly, this is the third VSCC-type-specific process found to change with age in long-term hippocampal culture. The other two include an increase in L-type VSCC density (Porter et al., 1997), and a substantial decline in membrane-delimited, G-protein-coupled inhibition of non-L-type VSCCs (Blalock et al., 1999). Thus, with age in culture, multiple changes in VSCC-type-specific regulation occur that seem likely to increase Ca2+ influx.

Although these VSCC alterations in culture may well reflect basic developmental processes (e.g., (Turrigiano et al., 1994), there also is some evidence that they may be, in part, analogous to brain aging processes. That is, changes in VSCC activity have been found consistently in models of brain aging and likely contribute to age-related alterations in Ca2+-dependent processes such as neuronal excitability, plasticity, and memory (cf., reviews, Khachaturian, 1989; Landfield et al., 1989; Disterhoft et al., 1993; Foster and Norris, 1997; Thibault et al., 1998; Verkhratsky and Toescu, 1998; Griffith et al., 2000). Further, growing evidence indicates that calcineurin function may be increased with brain aging (Norris et al., 1998; Foster et al., 2001) and, in preliminary studies, FK-506 inhibited VSCC current in hippocampal slice neurons of aged rats (Norris et al., 2001).

Additional studies will of course be required to determine whether similar aging changes in calcineurin content or sensitivity occur during aging in vivo. However, regardless of whether calcineurin regulation changes with brain aging, the present studies suggest that calcineurin plays a major enabling/enhancing role in normal and pathological processes mediated by L-type VSCC activity. As a corollary, modulation of calcineurin may provide a basis for developing new therapeutic strategies aimed at regulating L-type VSCC activity and L-type VSCC-dependent disorders.

Acknowledgments

This work was supported in part by Grants from the National Institute on Aging (AG04542, AG10836) and Training Grants (AG00242 and 1F32 AG05903). We wish to thank Elsie Barr, Janice Staton, and Veronique Thibault for important technical assistance, and Kelley Secrest and Judy Hower for excellent assistance with the manuscript.

References
  • Ankarcrona, M; Dypbukt, JM; Orrenius, S; Nicotera, P. Calcineurin and mitochondrial function in glutamate-induced neuronal cell death. FEBS Lett. 1996;394:321–324. [PubMed]
  • Armstrong, DL. Calcium channel regulation by calcineurin, a Ca2+-activated phosphatase in mammalian brain. Trends Neurosci. 1989;12:117–122. [PubMed]
  • Artalejo, CR; Rossie, S; Perlman, RL; Fox, AP. Voltage-dependent phosphorylation may recruit Ca2+ current facilitation in chromaffin cells. Nature. 1992;358:63–66. [PubMed]
  • Asai, A; Qiu, J; Narita, Y; Chi, S; Saito, N; Shinoura, N; Hamada, H; Kuchino, Y; Kirino, T. High level calcineurin activity predisposes neuronal cells to apoptosis. J Biol Chem. 1999;274:34450–34458. [PubMed]
  • Banker, GA; Cowan, WM. Rat hippocampal neurons in dispersed cell culture. Brain Res. 1977;126:337–342.
  • Bean, BP. Multiple types of calcium channels in heart muscle and neurons. Modulation by drugs and neurotransmitters. Ann N Y Acad Sci. 1989;560:334–345. [PubMed]
  • Bers, DM; Patton, CW; Nuccitelli, R. A practical guide to the preparation of Ca2+ buffers. Methods Cell Biol. 1994;40:3–29. [PubMed]
  • Birnbaumer, L; Campbell, KP; Catterall, WA; Harpold, MM; Hofmann, F; Horne, WA; Mori, Y; Schwartz, A; Snutch, TP; Tanabe, T, et al. The naming of voltage-gated calcium channels. Neuron. 1994;13:505–506. [PubMed]
  • Blalock, EM; Porter, NM; Landfield, PW. Decreased G-protein-mediated regulation and shift in calcium channel types with age in hippocampal cultures. J Neurosci. 1999;19:8674–8684. [PubMed]
  • Branchaw, JL; Banks, MI; Jackson, MB. Ca2+- and voltage-dependent inactivation of Ca2+ channels in nerve terminals of the neurohypophysis. J Neurosci. 1997;17:5772–5781. [PubMed]
  • Brewer, LD; Thibault, V; Chen, KC; Langub, MC; Landfield, PW; Porter, NM. Vitamin D hormone confers neuroprotection in parallel with downregulation of L-type calcium channel expression in hipocampal neurons. J Neurosci. 2001;21:98–108. [PubMed]
  • Brown, TH; Johnston, D. Voltage-clamp analysis of mossy fiber synaptic input to hippocampal neurons. J Neurophysiol. 1983;50:487–507. [PubMed]
  • Burley, JR; Sihra, TS. A modulatory role for protein phosphatase 2B (calcineurin) in the regulation of Ca2+ entry. Eur J Neurosci. 2000;12:2881–2891. [PubMed]
  • Cameron, AM; Steiner, JP; Roskams, AJ; Ali, SM; Ronnett, GV; Snyder, SH. Calcineurin associated with the inositol 1,4,5-trisphosphate receptor-FKBP12 complex modulates Ca2+ flux. Cell. 1995;83:463–472. [PubMed]
  • Campbell, LW; Hao, SY; Thibault, O; Blalock, EM; Landfield, PW. Aging changes in voltage-gated calcium currents in hippocampal CA1 neurons. J Neurosci. 1996;16:6286–6295. [PubMed]
  • Carafoli, E; Genazzani, A; Geurini, D. Calcium controls the transcription of its own transporters and channels in developing neurons. Biochem Biophys Res Commun. 1999;266:624–632. [PubMed]
  • Catterall, WA. Structure and function of voltage-gated ion channels. Annu Rev Biochem. 1995;64:493–531. [PubMed]
  • Chen, KC; Blalock, EM; Thibault, O; Kaminker, P; Landfield, PW. Expression of α1D subunit mRNA is correlated with L-type Ca2+ channel activity in single neurons of hippocampal ‘zipper’ slices. Proc Natl Acad Sci USA. 2000;97:4357–4362. [PubMed]
  • Church, PJ; Stanley, EF. Single L-type calcium channel conductance with physiological levels of calcium in chick ciliary ganglion neurons. J Physiol. 1996;496:59–68. [PubMed]
  • Cohen, P. Protein phosphorylation and hormone action. Proc R Soc Lond B Biol Sci. 1988;234:115–144. [PubMed]
  • Cohen, P. Classification of protein-serine/threonine phosphatases: identification and quantitation in cell extracts. Methods Enzymol. 1991;201:389–398. [PubMed]
  • Connern, CP; Halestrap, AP. Recruitment of mitochondrial cyclophilin to the mitochondrial inner membrane under conditions of oxidative stress that enhance the opening of a calcium-sensitive non-specific channel. Biochem J. 1994;302:321–324. [PubMed]
  • Corey, D.P., Stevens, C.F. (1983) Science and technology of patch recording electrodes. In: Sakmann B., Neher E. (Eds.), Single-channel Recording. Plenum, New York, pp. 53–58.
  • Cox, DH; Dunlap, K. Pharmacological discrimination of N-type from L-type calcium current and its selective modulation by transmitters. J Neurosci. 1992;12:906–914. [PubMed]
  • Currie, KP; Fox, AP. Comparison of N- and P/Q-type voltage-gated calcium channel current inhibition. J Neurosci. 1997;17:4570–4579. [PubMed]
  • Davare, MA; Horne, MC; Hell, JW. Protein phosphatase 2A is associated with class C L-type calcium channels (Cav1.2) and antagonizes channel phosphorylation by cAMP-dependent protein kinase. J Biol Chem. 2000;275:39710–39717. [PubMed]
  • Disterhoft, JF; Moyer, JR, Jr; Thompson, LT; Kowalska, M. Functional aspects of calcium-channel modulation. Clin Neuropharmacol. 1993;16:S12–S24. [PubMed]
  • Dolphin, AC. The G.L. Brown Prize Lecture. Voltage-dependent calcium channels and their modulation by neurotransmitters and G proteins. Exp Physiol. 1995;80:1–36. [PubMed]
  • Dolphin, AC. Facilitation of Ca2+ current in excitable cells. Trends Neurosci. 1996;19:35–43. [PubMed]
  • Dumont, FJ; Melino, MR; Staruch, MJ; Koprak, SL; Fischer, PA; Sigal, NH. The immunosuppressive macrolides FK-506 and rapamycin act as reciprocal antagonists in murine T cells. J Immunol. 1990;144:1418–1424. [PubMed]
  • Fisher, RE; Gray, R; Johnston, D. Properties and distribution of single voltage-gated calcium channels in adult hippocampal neurons. J Neurophysiol. 1990;64:91–104. [PubMed]
  • Foster, TC; Norris, CM. Age-associated changes in Ca2+-dependent processes: relation to hippocampal synaptic plasticity. Hippocampus. 1997;7:602–612. [PubMed]
  • Foster, TC; Sharrow, KM; Masse, JR; Norris, CM; Kumar, A. Calcineurin links Ca2+ dysregulation with brain aging. J Neurosci. 2001;21:4066–4073. [PubMed]
  • Fox, AP; Nowycky, MC; Tsien, RW. Single-channel recordings of three types of calcium channels in chick sensory neurones. J Physiol Lond. 1987;394:173–200. [PubMed]
  • Friberg, H; Ferrand-Drake, M; Bengtsson, F; Halestrap, AP; Wieloch, T. Cyclosporin A, but not FK 506, protects mitochondria and neurons against hypoglycemic damage and implicates the mitochondrial permeability transition in cell death. J Neurosci. 1998;18:5151–5159. [PubMed]
  • Genazzani, AA; Carafoli, E; Guerini, D. Calcineurin controls inositol 1,4,5-trisphosphate type 1 receptor expression in neurons. Proc Natl Acad Sci USA. 1999;96:5797–5801. [PubMed]
  • Graef, IA; Mermelstein, PG; Stankunas, K; Neilson, JR; Deisseroth, K; Tsien, RW; Crabtree, GR. L-type calcium channels and GSK-3 regulate the activity of NF-ATc4 in hippocampal neurons. Nature. 1999;401:703–708. [PubMed]
  • Griffith, WH; Jasek, MC; Bain, SH; Murchison, D. Modification of ion channels and calcium homeostasis of basal forebrain neurons during aging. Behav Brain Res. 2000;115:219–233. [PubMed]
  • Hamill, OP; Marty, A; Neher, E; Sakmann, B; Sigworth, FJ. Improved patch-clamp techniques for high-resolution current recording from cells and cell-free membrane patches. Pflug Arch. 1981;391:85–100.
  • Hashimoto, Y; Perrino, BA; Soderling, TR. Identification of an autoinhibitory domain in calcineurin. J Biol Chem. 1990;265:1924–1927. [PubMed]
  • Hell, JW; Westenbroek, RE; Breeze, LJ; Wang, KK; Chavkin, C; Catterall, WA. N-methyl-d-aspartate receptor-induced proteolytic conversion of postsynaptic class C L-type calcium channels in hippocampal neurons. Proc Natl Acad Sci USA. 1996;93:3362–3367. [PubMed]
  • Hell, JW; Westenbroek, RE; Warner, C; Ahlijanian, MK; Prystay, W; Gilbert, MM; Snutch, TP; Catterall, WA. Identification and differential subcellular localization of the neuronal class C and class D L-type calcium channel α1 subunits. J Cell Biol. 1993a;123:949–962. [PubMed]
  • Hell, JW; Yokoyama, CT; Wong, ST; Warner, C; Snutch, TP; Catterall, WA. Differential phosphorylation of two size forms of the neuronal class C L-type calcium channel α1 subunit. J Biol Chem. 1993b;268:19451–19457. [PubMed]
  • Herman, JP; Chen, KC; Booze, R; Landfield, PW. Up-regulation of α1D Ca2+ channel subunit mRNA expression in the hippocampus of aged F344 rats. Neurobiol Aging. 1998;19:581–587. [PubMed]
  • Hernandez-Lopez, S; Tkatch, T; Perez-Garci, E; Galarraga, E; Bargas, J; Hamm, H; Surmeier, DJ. D2 dopamine receptors in striatal medium spiny neurons reduce L-type Ca2+ currents and excitability via a novel PLCβ1-IP3-calcineurin-signaling cascade. J Neurosci. 2000;20:8987–8995. [PubMed]
  • Hescheler, J; Rosenthal, W; Trautwein, W; Schultz, G. The GTP-binding protein, Go, regulates neuronal calcium channels. Nature. 1987;325:445–447. [PubMed]
  • Hille, B. Modulation of ion-channel function by G-protein-coupled receptors. Trends Neurosci. 1994;17:531–536. [PubMed]
  • Holz, GGT; Rane, SG; Dunlap, K. GTP-binding proteins mediate transmitter inhibition of voltage-dependent calcium channels. Nature. 1986;319:670–672. [PubMed]
  • Ikeda, SR; Dunlap, K. Voltage-dependent modulation of N-type calcium channels: role of G protein subunits. Adv Second Messenger Phosphoprotein Res. 1999;33:131–151. [PubMed]
  • Jiang, LH; Gawler, DJ; Hodson, N; Milligan, CJ; Pearson, HA; Porter, V; Wray, D. Regulation of cloned cardiac L-type calcium channels by cGMP-dependent protein kinase. J Biol Chem. 2000;275:6135–6143. [PubMed]
  • Johnston, D; Brown, TH. Interpretation of voltage-clamp measurements in hippocampal neurons. J Neurophysiol. 1983;50:464–486. [PubMed]
  • Jones, MV; Westbrook, GL. Shaping of IPSCs by endogenous calcineurin activity. J Neurosci. 1997;17:7626–7633. [PubMed]
  • Khachaturian, ZS. The role of calcium regulation in brain aging: reexamination of a hypothesis. Aging Milano. 1989;1:17–34. [PubMed]
  • Krupp, JJ; Vissel, B; Heinemann, SF; Westbrook, GL. Calcium-dependent inactivation of recombinant N-methyl-d-aspartate receptors is NR2 subunit specific. Mol Pharmacol. 1996;50:1680–1688. [PubMed]
  • Kung, L; Halloran, PF. Immunophilins may limit calcineurin inhibition by cyclosporine and tacrolimus at high drug concentrations. Transplantation. 2000;70:327–335. [PubMed]
  • Kuno, T; Mukai, H; Ito, A; Chang, CD; Kishima, K; Saito, N; Tanaka, C. Distinct cellular expression of calcineurin Aα and Aβ in rat brain. J Neurochem. 1992;58:1643–1651. [PubMed]
  • Landfield, PW; Campbell, LW; Hao, SY; Kerr, DS. Aging-related increases in voltage-sensitive, inactivating calcium currents in rat hippocampus. Implications for mechanisms of brain aging and Alzheimer’s disease. Ann N Y Acad Sci. 1989;568:95–105. [PubMed]
  • Legendre, P; Rosenmund, C; Westbrook, GL. Inactivation of NMDA channels in cultured hippocampal neurons by intracellular calcium. J Neurosci. 1993;13:674–684. [PubMed]
  • Liu, J; Farmer, JD, Jr; Lane, WS; Friedman, J; Weissman, I; Schreiber, SL. Calcineurin is a common target of cyclophilin-cyclosporin A and FKBP-FK506 complexes. Cell. 1991;66:807–815. [PubMed]
  • Llinas, R; Sugimori, M; Hillman, DE; Cherksey, B. Distribution and functional significance of the P-type, voltage-dependent Ca2+ channels in the mammalian central nervous system. Trends Neurosci. 1992;15:351–355. [PubMed]
  • Lukyanetz, EA. Evidence for colocalization of calcineurin and calcium channels in dorsal root ganglion neurons. Neuroscience. 1997;78:625–628. [PubMed]
  • Lukyanetz, EA; Piper, TP; Sihra, TS. Calcineurin involvement in the regulation of high-threshold Ca2+ channels in NG108-15 (rodent neuroblastoma×glioma hybrid) cells. J Physiol Lond. 1998;510:371–385. [PubMed]
  • Mansuy, IM; Mayford, M; Jacob, B; Kandel, ER; Bach, ME. Restricted and regulated overexpression reveals calcineurin as a key component in the transition from short-term to long-term memory. Cell. 1998;92:39–49. [PubMed]
  • Marunaka, Y; Niisato, N; Shintani, Y. Protein phosphatase 2B-dependent pathway of insulin action on single Cl-channel conductance in renal epithelium. J Membr Biol. 1998;161:235–245. [PubMed]
  • McDonough, SI; Swartz, KJ; Mintz, IM; Boland, LM; Bean, BP. Inhibition of calcium channels in rat central and peripheral neurons by omega-conotoxin MVIIC. J Neurosci. 1996;16:2612–2623. [PubMed]
  • Mintz, IM; Bean, BP. Block of calcium channels in rat neurons by synthetic omega-Aga-IVA. Neuropharmacology. 1993;32:1161–1169. [PubMed]
  • Norris, CM; Blalock, EM; Landfield, PW. Inhibition of calcineurin reduces Ca2+ channel currents in hippocampal neurons: enhanced effect with age-in-culture. Soc Neurosci Abstr. 1999;25:326–331.
  • Norris, CM; Chen, KC; Landfield, PW. Inhibition of calcineurin reduces voltage-gated calcium channel currents in CA1 neurons of hippocampal ‘zipper’ slices. Soc Neurosci Abstr. 2001;27:100–101.
  • Norris, CM; Halpain, S; Foster, TC. Alterations in the balance of protein kinase/phosphatase activities parallel reduced synaptic strength during aging. J Neurophysiol. 1998;80:1567–1570. [PubMed]
  • Perrino, BA; Ng, LY; Soderling, TR. Calcium regulation of calcineurin phosphatase activity by its B subunit and calmodulin. Role of the autoinhibitory domain (published erratum appears in J Biol Chem 270 (1995) 7012). J Biol Chem. 1995;270:340–346. [PubMed]
  • Polli, JW; Billingsley, ML; Kincaid, RL. Expression of the calmodulin-dependent protein phosphatase, calcineurin, in rat brain: developmental patterns and the role of nigrostriatal innervation. Brain Res Dev Brain Res. 1991;63:105–119.
  • Porter, NM; Thibault, O; Thibault, V; Chen, KC; Landfield, PW. Calcium channel density and hippocampal cell death with age in long-term culture. J Neurosci. 1997;17:5629–5639. [PubMed]
  • Randall, A; Tsien, RW. Pharmacological dissection of multiple types of Ca2+ channel currents in rat cerebellar granule neurons. J Neurosci. 1995;15:2995–3012. [PubMed]
  • Sagoo, JK; Fruman, DA; Wesselborg, S; Walsh, CT; Bierer, BE. Competitive inhibition of calcineurin phosphatase activity by its autoinhibitory domain. Biochem J. 1996;320:879–884. [PubMed]
  • Sakmann, B., Neher, E., 1983. Geometric parameters of pipettes and membrane patches. In: Sakmann, B., Neher E. (Eds.), Single Channel Recording. Plenum, New York, pp. 37–51.
  • Schuhmann, K; Romanin, C; Baumgartner, W; Groschner, K. Intracellular Ca2+ inhibits smooth muscle L-type Ca2+ channels by activation of protein phosphatase type 2B and by direct interaction with the channel. J Gen Physiol. 1997;110:503–513. [PubMed]
  • Snyder, SH; Lai, MM; Burnett, PE. Immunophilins in the nervous system. Neuron. 1998;21:283–294. [PubMed]
  • Spruston, N; Jaffe, DB; Johnston, D. Dendritic attenuation of synaptic potentials and currents: the role of passive membrane properties. Trends Neurosci. 1994;17:161–166. [PubMed]
  • Stemmer, PM; Klee, CB. Dual calcium ion regulation of calcineurin by calmodulin and calcineurin B. Biochemistry. 1994;33:6859–6866. [PubMed]
  • Stewart, AA; Ingebritsen, TS; Manalan, A; Klee, CB; Cohen, P. Discovery of a Ca2+- and calmodulin-dependent protein phosphatase: probable identity with calcineurin (CaM-BP80). FEBS Lett. 1982;137:80–84. [PubMed]
  • Surmeier, DJ; Bargas, J; Hemmings, HC, Jr; Nairn, AC; Greengard, P. Modulation of calcium currents by a D1 dopaminergic protein kinase/phosphatase cascade in rat neostriatal neurons. Neuron. 1995;14:385–397. [PubMed]
  • Thibault, O; Landfield, PW. Increase in single L-type calcium channels in hippocampal neurons during aging. Science. 1996;272:1017–1020. [PubMed]
  • Thibault, O; Porter, NM; Chen, KC; Blalock, EM; Kaminker, PG; Clodfelter, GV; Brewer, LD; Landfield, PW. Calcium dysregulation in neuronal aging and Alzheimer’s disease: history and new directions. Cell Calcium. 1998;24:417–433. [PubMed]
  • Thibault, O; Porter, NM; Landfield, PW. Low Ba2+ and Ca2+ induce a sustained high probability of repolarization openings of L-type Ca2+ channels in hippocampal neurons: physiological implications. Proc Natl Acad Sci USA. 1993;90:11792–11796. [PubMed]
  • Tong, G; Shepherd, D; Jahr, CE. Synaptic desensitization of NMDA receptors by calcineurin. Science. 1995;267:1510–1512. [PubMed]
  • Tsien, RW; Lipscombe, D; Madison, D; Bley, K; Fox, A. Reflections on Ca2+-channel diversity, 1988–1994. Trends Neurosci. 1995;18:52–54. [PubMed]
  • Turrigiano, G; Abbott, LF; Marder, E. Activity-dependent changes in the intrinsic properties of cultured neurons. Science. 1994;264:974–977. [PubMed]
  • Verkhratsky, A; Toescu, EC. Calcium and neuronal ageing. Trends Neurosci. 1998;21:2–7. [PubMed]
  • Victor, RG; Rusnak, F; Sikkink, R; Marban, E; O’Rourke, B. Mechanism of Ca2+-dependent inactivation of L-type Ca2+ channels in GH3 cells: direct evidence against dephosphorylation by calcineurin. J Membr Biol. 1997;156:53–61. [PubMed]
  • Wang, HG; Pathan, N; Ethell, IM; Krajewski, S; Yamaguchi, Y; Shibasaki, F; McKeon, F; Bobo, T; Franke, TF; Reed, JC. Ca2+-induced apoptosis through calcineurin dephosphorylation of BAD. Science. 1999;284:339–343. [PubMed]
  • Westenbroek, RE; Hell, JW; Warner, C; Dubel, SJ; Snutch, TP; Catterall, WA. Biochemical properties and subcellular distribution of an N-type calcium channel alpha 1 subunit. Neuron. 1992;9:1099–1115. [PubMed]
  • Winder, DG; Mansuy, IM; Osman, M; Moallem, TM; Kandel, ER. Genetic and pharmacological evidence for a novel, intermediate phase of long-term potentiation suppressed by calcineurin. Cell. 1998;92:25–37. [PubMed]
  • Yasutsune, T; Kawakami, N; Hirano, K; Nishimura, J; Yasui, H; Kitamura, K; Kanaide, H. Vasorelaxation and inhibition of the voltage-operated Ca2+ channels by FK506 in the porcine coronary artery. Br J Pharmacol. 1999;126:717–729. [PubMed]
  • Yatani, A; Honda, R; Tymitz, KM; Lalli, MJ; Molkentin, JD. Enhanced Ca2+ channel currents in cardiac hypertrophy induced by activation of calcineurin-dependent pathway. J Mol Cell Cardiol. 2001;33:249–259. [PubMed]
  • Zamponi, GW; Snutch, TP. Modulation of voltage-dependent calcium channels by G proteins. Curr Opin Neurobiol. 1998;8:351–356. [PubMed]
  • Zeilhofer, HU; Blank, NM; Neuhuber, WL; Swandulla, D. Calcium-dependent inactivation of neuronal calcium channel currents is independent of calcineurin. Neuroscience. 2000;95:235–241. [PubMed]
  • Zhu, Y; Yakel, JL. Calcineurin modulates G protein-mediated inhibition of N-type calcium channels in rat sympathetic neurons. J Neurophysiol. 1997;78:1161–1165. [PubMed]
  • Zong, X; Schreieck, J; Mehrke, G; Welling, A; Schuster, A; Bosse, E; Flockerzi, V; Hofmann, F. On the regulation of the expressed L-type calcium channel by cAMP-dependent phosphorylation. Pflug Arch. 1995;430:340–347.