Skip to main content
Biophysical Journal logoLink to Biophysical Journal
. 2007 May 25;93(4):1175–1183. doi: 10.1529/biophysj.106.103333

Zn2+ Sensitivity of High- and Low-Voltage Activated Calcium Channels

Hong-Shuo Sun 1, Kwokyin Hui 1, David W K Lee 1, Zhong-Ping Feng 1
PMCID: PMC1929049  PMID: 17526568

Abstract

The essential cation zinc (Zn2+) blocks voltage-dependent calcium channels in several cell types, which exhibit different sensitivities to Zn2+. The specificity of the Zn2+ effect on voltage-dependent calcium channel subtypes has not been systematically investigated. In this study, we used a transient protein expression system to determine the Zn2+ effect on low- and high-voltage activated channels. We found that in Ba2+, the IC50 value of Zn2+ was α1-subunit-dependent with lowest value for CaV1.2, and highest for CaV3.1; the sensitivity of the channels to Zn2+ was approximately ranked as CaV1.2 > CaV3.2 > CaV2.3 > CaV2.2 = CaV 2.1 ≥ CaV3.3 = CaV3.1. Although the CaV2.2 and CaV3.1 channels had similar IC50 for Zn2+ in Ba2+, the CaV2.2, but not CaV3.1 channels, had ∼10-fold higher IC50 to Zn2+ in Ca2+. The reduced sensitivity of CaV2.2 channels to Zn2+ in Ca2+ was partially reversed by disrupting a putative EF-hand motif located external to the selectivity filter EEEE locus. Thus, our findings support the notion that the Zn2+ block, mediated by multiple mechanisms, may depend on conformational changes surrounding the α1 pore regions. These findings provide fundamental insights into the mechanism underlying the inhibitory effect of zinc on various Ca2+ channel subtypes.

INTRODUCTION

The regulation of calcium entry into cells via voltage-dependent calcium channels (VDCCs) plays a fundamental role in controlling synaptic transmission, membrane excitability, muscle contraction, rhythmic activity, gene transcription, and signal transduction pathways (1,2). Therefore, elucidating the molecular mechanisms regulating calcium channel conductivity is essential for a greater comprehension of cell biology.

Zinc, an essential transition divalent cation, is involved in maintaining and regulating cellular and subcellular functions of virtually all cells. Four major roles of Zn2+ have been reported: 1), it binds tightly to metalloenzymes and serves as a cofactor in gene expression and enzymatic reactions (3); 2), it is abundant in the brain, where it is localized in the presynaptic terminals, is coreleased with glutamate (4,5), and regulates neuronal excitability and synaptic plasticity (6,7); 3), it is involved in many pathological neuromodulatory events, and the accumulation of Zn2+ during and after transient global ischemia (8,9) can also affect cardiac functions (10,11); and 4), it regulates conductivity of VDCCs (1214) and affects Ca2+ signaling.

Zn2+ block of VDCCs has been found in various neuronal preparations from different species, including DRG neurons (12,15), hypothalamic neurons (13,16), paleocortical neurons (17), thalamic relay neurons (14), and pelvic neurons (18). The sensitivity of native calcium channels to Zn2+ is highly variable, with IC50 values ranging from 7 μM (13) to 300 μM (15). In these studies, Zn2+ inhibition of the VDCCs was primarily affected by the animal species used and the cell type studied (14,19). Most neurons express multiple calcium channel subtypes, and each channel has a distinct physiological role. Thus, the sensitivity of specific calcium channels to zinc inhibition, which has not yet been determined, is critically important for a better understanding of their particular role in calcium conductance.

Inhibition of different calcium channel subtypes by various inorganic and organic calcium channel blockers (20,21) has been studied in cells in which the channel subtypes were transiently expressed. High-voltage–activated (HVA) calcium channels are heteromultimers consisting of α1, α2, β-, and δ-subunits. The α1 subunit, which contains the channel pore region, encompassing the selectivity filter locus, plays the predominant role in determining channel conductance. Currently, 11 VDCC α1 subunits have been cloned, including four HVA CaV1 L-type channels (CaV1.1/α1S, CaV1.2/α1C, CaV1.3/α1D, and CaV1.4/α1F); three HVA CaV2 non-L-type calcium channels (CaV2.1/α1A, CaV2.2/α1B, and CaV2.3/α1E); and three low-voltage-activated (LVA) CaV3 T-type channels (CaV3.1/α1G, CaV3.2/α1H, and CaV3.3/α1I) (2224). The electrophysiological and pharmacological properties of the α1 subunits have been well described. For instance, α1C (CaV1.2) is found in dihydropyridine-sensitive L-type channels; α1A (CaV2.1) in ω-agatoxin IVA-sensitive P-/Q-type channels; α1B (CaV2.2) in ω-conotoxin GVIA-sensitive N-type channels; and α1E (CaV2.3) in R-type channels. T-type calcium channels contain one of three α1 subunits, α1G (CaV3.1), α1H (CaV3.2), and α1I (CaV3.3), and their specific pharmacological properties have not yet been clearly identified (25). In this study, we systematically determined the effect of Zn2+ on transiently expressed HVA and LVA calcium channels and demonstrated that sensitivity to Zn2+ inhibition is α1 subtype specific and may be dependent on selectivity filter residues outside of the EEEE/EEDD locus of the channels.

METHODS

Tissue culture and transient transfection

Human embryonic kidney tsA-201 cells were maintained in standard DMEM supplemented with 10% fetal bovine serum, 200 units/ml penicillin, and 0.2 mg/ml streptomycin at 37°C in a CO2 incubator (26,27). Cells were split with trypsin-EDTA, plated on glass coverslips at 10% confluency, and allowed to recover for 12 h at 37°C. The cells were then transiently transfected with expression vectors containing cDNAs encoding wild-type or mutant calcium channel α1, β1b, and α2-δ subunits and enhanced green fluorescent protein (eGFP) at a 1:1:1:0.2 molar ratio, using a standard Ca2+ phosphate protocol (26). All wild-type α1 (α1G, AF290212; α1H, AF290213; α1I, AF290214; α1C, M67515; α1A, M64373; α1B, M92905; and α1E, L15453), β1b, and α2-δ subunits cDNAs were generous gifts from Dr. Terry Snutch (University of British Columbia, Vancouver, British Columbia, Canada), and mutant α1B subunit cDNAs were generous gifts from Dr. Gerald Zamponi (University of Calgary, Calgary, Alberta, Canada). After 12 h, the cell culture medium was replaced with fresh medium. The cells were allowed to recover for an additional 12 h and were subsequently kept at 28°C in 5% CO2 for 1–2 days before physiological recordings were made.

Chemicals and solutions

All chemicals used in the cell culture were purchased from GIBCO (Invitrogen, Burlington, Canada). Chemicals used for physiological recordings were purchased from Sigma (St. Louis, MO).

Electrophysiology

Whole-cell patch-clamp (ruptured) recordings were performed using a MultiClamp 700A amplifier (Axon Instruments, Foster City, CA) linked to a personal computer equipped with pClamp9. Patch pipettes (Sutter borosilicate glass, BF 150-86-15) were pulled using a Sutter P-87 microelectrode puller and subsequently fire polished using a Narashige microforge. Pipettes (in the range of 2–4 MΩ) were filled with internal solution containing 108 mM Cs-methanesulfonate, 4 mM MgCl2, 9 mM EGTA, and 9 mM HEPES (adjusted to pH 7.2 with TEA-OH). The cells were transferred to a 3.5-cm culture dish containing recording solution comprised of 20 mM BaCl2 (or CaCl2), 1 mM MgCl2, 10 mM HEPES, 40 mM TEACl, 10 mM glucose, and 87.5 mM CsCl (adjusted to pH 7.2 with TEA-OH). Currents were elicited by stepping from a holding potential of −100 mV to various test potentials; Clampex software was used to control this process. Data were filtered at 1 kHz using a four-pole Bessel filter and digitized at a sampling frequency of 2 kHz.

Data analyses

All data were analyzed using Clampfit (Axon); curve fittings were carried out using SigmaPlot 4.0 (Jandel Scientific). Dose-response curves were fitted using the equation I/I0 = 1/(1 + ([Zn2+]/IC50)n), where I is the peak current response to a given test potential in the presence of Zn2+, I0 is the current obtained in the drug-free condition (the control condition), [Zn2+] is the zinc concentration, IC50 is the concentration at which 50% inhibition is obtained, and n is the Hill coefficient. Current-voltage relations obtained from peak current amplitude were fitted to the equation I = (Gmax (VVr))/(1 + exp((VhV)/S)), where I is the measured peak current, Gmax is the slope conductance, V is the test potential, Vr is the apparent reversal potential, Vh is the potential of half-maximal activation, and S is the slope of activation.

Statistics

Statistical analyses were carried out using SigmaStat 3.0 software (SPSS, Chicago, IL). The significance of differences between mean values from each experimental group was tested using a Student's t-test for two groups and one-way analysis of variance for multiple comparisons. Differences were considered significant if p < 0.05.

RESULTS AND DISCUSSION

Sensitivities of calcium channels to zinc inhibition are α1 subunit specific

Differential inhibitory effects of Zn2+ on LVA calcium channels

We investigated the inhibitory effect of Zn2+ on three recombinant low-voltage–dependent T-type calcium channels (α1G, α1H, and α1I) using a transient expression system. We evoked whole-cell currents in tsA-201 cells transiently expressing T-type calcium channels by a 10-ms depolarization step of −20 mV from a holding potential of −100 mV in 20 mM Ba2+. Representative current recordings for cells expressing CaV3.2 (α1H) obtained for various Zn2+ concentrations are shown in Fig. 1 A1, and the corresponding time course of the development of the Zn2+ inhibition on the tail current is shown in Fig. 1 A2. Inhibition of the current was observed within 10 s of exposing cells to Zn2+ and rapidly reached equilibrium. Inhibition of the T-type calcium channel was reversible, as rapid recovery from Zn2+ inhibition was nearly complete after washout and removal of the Zn2+, consistent with previous reports for native neurons (14,19). Fig. 1 B shows a mean concentration-response curve fitted to the Hill equation (with a Hill coefficient of ∼1). As determined from eight independent experiments, the mean IC50 value of Zn2+ for the CaV3.2 channel was 24.1 ± 1.9 μM.

FIGURE 1.

FIGURE 1

Differential block effect of Zn2+ on LVA calcium channels. (A) A typical representative of the dose-dependent blockade effect of Zn2+ on whole-cell current of CaV3.2 (α1H) calcium channels. (A1) Current traces recorded in 20 mM Ba2+ for CaV3.2 channels during 10-ms voltage steps from a holding potential of −100 mV to −20 mV in the absence or presence of various Zn2+ concentrations, as indicated. Note that the current amplitude decreased with an increase in Zn2+ concentrations. (A2) The corresponding time course of development of block and recovery from Zn block in CaV3.2 channels is shown in A1. Comparison of the average dose-response curves (B) and IC50 values (C) of Zn2+ block effect on three LVA calcium channels. The cells have been recorded from: CaV3.1 (α1G), n = 7; CaV3.2 (α1H), n = 8; and CaV3.3 (α1I), n = 7. The data are presented as mean ± SE. (*) Statistical significance (p < 0.05) among the groups. The number in the parentheses in C indicates the number of cells used for the recordings.

The inhibitory effect of Zn2+ on CaV3.1 (α1G) and CaV3.3 (α1I) channels was determined using the same protocol as was used for the CaV3.2 channels (α1H). Interestingly, these channels exhibited a different sensitivity to Zn2+, as the mean IC50 values for Zn2+ varied depending on the α1 subunit contained in the channel. Fig. 1 B shows a comparison of the mean concentration-response curves of CaV3.2 (α1H) and CaV3.3 (α1I) channels to the response curves for CaV3.1 (α1G) channels. The mean concentration-response curve of the CaV3.2 channel shifted to a lower concentration than the response curves of the CaV3.1 or CaV3.3 channels (Fig. 1 B), indicating that the CaV3.2 channels are more sensitive to Zn2+ inhibition than are the other T-type calcium channels. The mean IC50 values of Zn2+ for the three LVA channels are compared in Fig. 1 C. The IC50 of Zn2+ for the CaV3.2 (α1H) channels (24.1 ± 1.9 μM, n = 8) was significantly lower than that of the CaV3.1 (α1G) (196.5 ± 50.4 μM, n = 7; p < 0.05) and CaV3.3 (α1I) channel (152.2 ± 30.6 μM, n = 8; p < 0.05). The Hill coefficient for CaV3.2 (α1H) was 1.3 ± 0.2, that for CaV3.2 (α1G) was 1.1 ± 0.1, and the one for CaV3.3 (α1I) was 1.0 ± 0.1. The differences of the Hill coefficients among three T-type channels were not statistically significant (p > 0.05), consistent with a recent report (28). These results demonstrate that the sensitivity to Zn2+ of the LVA calcium channels is intrinsically different among the subtypes, with CaV3.2-type channels being the most sensitive to the inhibitory effects of Zn2+. These findings are consistent with previous reports that α1H has a higher sensitivity to Zn2+ than α1G and α1I (18,28). The Hill coefficients of Zn2+ for T-type channels varied ∼1, indicating a single binding site or no cooperative interaction between Zn2+ binding sites under the Ba2+ condition. The change in the potency of Zn2+ among the LVA channels likely results from the different binding affinities, assuming that Zn2+ binds to a similar site. Because the selectivity filter of all T-type channels has a conserved EEDD motif (2931), the differing sensitivities, reflected by IC50 values, of the T-type channels indicate that the Zn2+ inhibitory effect on these channels is likely regulated by the residues outside of the pore EEDD motif.

Differential inhibitory effects of Zn2+ on HVA calcium channels

We next performed whole-cell current recordings to determine the effect of Zn2+ on recombinant high- and intermediate-voltage–dependent calcium channels transiently expressed in tsA-201 cells. Similar to the results obtained for the LVA calcium channels, we found that Zn2+ inhibited the HVA calcium channels in a concentration-dependent manner (Fig. 2). Representative Ba2+ current recordings obtained for the L-type calcium channels (α1C + β1b + α2δ) in various concentrations of Zn2+ are shown in Fig. 2 A, and the corresponding time course of the development for Zn2+ inhibition is shown in Fig. 2 B. We evoked currents by using a depolarization step of +10 mV from a holding potential of −100 mV in 20 mM Ba2+. Inhibition of the current developed within 10 s of exposure to Zn2+, rapidly reached equilibrium, and was eliminated promptly after washout of the Zn2+ (Fig. 2 A2), similar to the results obtained in experiments using transiently expressed T-type calcium channels (Fig. 1 B). Fig. 2 B shows the mean concentration-response curve fitted using the Hill equation (with a Hill coefficient of ∼1). As determined from eight independent experiments, the mean IC50 value of Zn2+ inhibition of the L-type channel was 10.9 ± 3.4 μM.

FIGURE 2.

FIGURE 2

Differential block effect of Zn2+ on HVA calcium channels. (A) A typical representative of the dose-dependent blockade effect of Zn2+ on whole-cell current of CaV1.2 (α1C) calcium channels. (A1) Current traces recorded in 20 mM Ba2+ for CaV1.2 channels during 10-ms voltage steps from a holding potential of −100 mV to −20 mV in the absence or presence of various Zn2+ concentrations, as indicated. Note that the current amplitude decreased with an increase in Zn2+ concentrations. (A2) The corresponding time course of development of block and recovery from Zn block in CaV1.2 channels is shown in A1. Comparison of the average dose-response curves (B) and IC50 values (C) of Zn2+ block effect on four HVA calcium channels. The Hill coefficients to Zn2+ were CaV1.2 (α1C), 0.88 ± 0.08 (n = 8); CaV2.1 (α1A), 0.91 ± 0.06 (n = 8); CaV2.2 (α1B), 0.92 ± 0.13 (n = 12); and CaV2.3 (α1E), 0.55 ± 0.05 (n = 8). The data are presented as mean ± SE. Statistical significance (p < 0.05) (*) to CaV1.2 (α1C) and (#) to CaV2.3 (α1E) when one-way analysis of variance was used for comparison. The number in the parentheses in C indicates the number of cells used for the recordings.

The inhibitory effects of Zn2+ on N- (α1B), P/Q- (α1A), and R-type (α1E) calcium channels (coexpressed with β1b and α2δ subunits) were determined using the same protocol used to determine the effect for L-type calcium channels (α1C). Similar to T-type channels, HVA calcium channels responded differentially to the inhibitory effect of zinc, as the mean IC50 values varied depending on the α1 subunit. As depicted in the concentration-response curves in Fig. 2 B, the mean concentration-response curves for N- (α1B), P/Q- (α1A), and R-type (α1E) calcium channels were shifted to higher concentrations compared to the response curve for the L-type calcium channel (Fig. 2 B), indicating that these channels are less sensitive to Zn2+ inhibition than is the L-type calcium channel. The IC50 values of all four HVA calcium channels are compared in Fig. 2 C. The Zn2+ IC50 value for the N-type (98.0 ± 17.9 μM, n = 12) and P/Q-type (110.0 ± 7.0 μM, n = 8) channels are ∼10-fold greater than that of the L-type (10.9 ± 3.4 μM, n = 8) channel. These differences are statistically significant (p < 0.05, Fig. 2 C). The Zn2+ IC50 value for the R-type calcium channel was 31.8 ± 12.3 μM (n = 8), which is significantly higher than that of the L-type channel (p < 0.05) but lower than that of N- and P-/Q-type channels (p < 0.05). These results clearly demonstrate that the sensitivity to the Zn2+ inhibitory effect of the HVA calcium channels differs intrinsically with the highest sensitivity found in the L-type channels and the lowest sensitivity in the N- and P/Q- type channels. Again, as seen in LVA calcium channels, the Zn2+ effect on HVA calcium channels varied depending on the channel α1 subtype, and this variation in sensitivity did not seem directly related to the selectivity filter EEEE residues, which are conserved in all HVA calcium channels (32,33).

Zn2+ block and the selectivity filter of calcium channels

The selectivity filter of the pore region of T-type channels has an EEDD locus (2931), whereas the selectivity filter in high-voltage–gated channels has an EEEE locus (32,33). We found that the variation in sensitivity to Zn2+ block of the channels was independent of these conserved selectivity filter residues. Based on the observed Zn2+ IC50 values for the various channels, we found that the order of sensitivity to the block by Zn2+ was approximately α1C > α1H > α1E > α1B = α1A > α1I = α1G when Ba2+ was used as the charge carrier.

Most HVA calcium channels conduct Ba2+ ∼2-fold better than they conduct Ca2+ (see McDonald et al. (34)). This difference seems to result from a higher binding affinity for Ca2+ ions than for Ba2+ ions at the EEEE locus of the channel pore region (26,32,33,35). In contrast, the Ca2+ conductance of the LVA channels is similar to or slightly greater than that of Ba2+ (3638). If Zn2+ inhibition of these channels depends on the charge carrier species and hence on the properties of the permeation pathway lining the pore region, the sensitivity of the HVA calcium channels to Zn2+ would be expected to be less in Ca2+ than in Ba2+, whereas that of T-type channels would remain unchanged. Therefore, we tested whether Zn2+ inhibition of these channels is affected by the charge carrier. Specifically, CaV2.2 N-type and CaV3.1 T-type calcium channels were compared because their IC50 values in 20 mM Ba2+ were similar (100–200 μM), when the inhibitory effect of Zn2+ on Ba2+ and Ca2+ currents was measured. Fig. 3 A shows the I-V curves for both channels in Ba2+ (Fig. 3, A1 and A3) and Ca2+ (Fig. 3, A2 and A4) with or without 100 μM Zn2+. Consistent with previous reports (26,3739), the current amplitude ratio between Ba2+ and Ca2+ (IBa/ICa) was ∼2 for the N-type channel and ∼0.8 for the CaV3.1 channel (Fig. 3, inset), and the I-V relations of the channels were not affected by Zn2+. As shown in Fig. 3 B, the IC50 for CaV2.2 α1B significantly increased (by ∼14-fold, p < 0.05) in 20 mM Ca2+ (1.21 ± 0.08 mM, n = 8) as compared to that in 20 mM Ba2+ (98.0 ± 17.9 μM, n = 12), suggesting that the current carrier species plays a role in regulating Zn2+ inhibition of the N-type channel. In contrast, the T-type CaV3.1 (α1G) channel had a similar sensitivity to Zn2+ in Ba2+ (IC50 198.5 ± 50.4 μM, n = 7) and in Ca2+ (IC50 132.8 ± 21.2 μM, n = 7, consistent with Jeong et al. (18)); and the difference in the IC50 values was not statistically significant (p > 0.05).

FIGURE 3.

FIGURE 3

Comparison of Zn effects on ICa and IBa of CaV2.2 (α1B) N-type and CaV3.1 (α1G) T-type calcium channels. (A) Representative current-voltage curves of CaV2.2 (α1B) N-type and CaV3.1 (α1G) T-type calcium channels in the absence or presence of 100 μM Zn2+: CaV2.2 (α1B) channels (A1); 20 mM Ba2+ (A2); 20 mM Ca2+. CaV3.1 (α1G) channels: (A3) 20 mM Ba2+; (A4) 20 mM Ca2+. (Inset) Comparison of the peak Ba2+ to Ca2+ current ratio of CaV2.2 (α1B) and CaV3.1 (α1G) calcium channels. (B) Summary of IC50 values of Zn2+ on CaV2.2 (α1B) and CaV3.1 (α1G) channels. The blockade effects of Zn2+ on CaV2.2 (α1B) channels are ∼10-fold higher in Ca2+ than in Ba2+ conditions, whereas those on CaV3.1 (α1G) channels are similar. The data are presented as mean ± SE. from 12 cells expressing CaV2.2 (α1B) channels and 6 cells expressing CaV3.1 (α1G) channels. (*) Statistical significance (p < 0.05) between the different conditions. The number in parentheses indicates the number of cells used for the recordings.

Multiple mechanisms of Zn2+ block of VDCCs have been suggested, including pore block, surface charge screening, and gating modification (14,19,40); however, these mechanisms vary among the different channel subtypes. Our findings show that the charge carrier–dependent Zn2+ sensitivity cannot be explained by either hydration energy or surface charge screening alone. Experimental hydration energies for Ba2+, Ca2+, and Zn2+ are 298, 360, and 467 kcal/mol (41,42), respectively. If ion access to the pore required complete removal of the hydration shells, then these energies indicate that permeation/block follows the order Ba2+ > Ca2+ > Zn2+. Thus, Zn2+ block would be expected to be more sensitive to Ba2+ than to Ca2+. Although this seems to be the case for the CaV2.2 (α1B) channel, it does not explain the differences seen in all the channels. For instance, all three T-type channels have similar Ba2+ to Ca2+ current ratios, ∼0.8, and the differences in Zn2+ sensitivities in the presence of Ca2+ and Ba2+ conditions are inconsistent. α1I sensitivity was substantially higher in Ba2+ (IC50 152 μM; Fig. 1 D) than reported in Ca2+ (IC50 470 μM (18)); α1H was lower in Ba2+ (IC50 ∼ 24 μM; Fig. 1 D) than in Ca2+ (IC50 2.4 μM (18)); and in this study we showed that Zn2+ sensitivity of CaV3.1 (α1G) was similar in Ba2+ and Ca2+ (Fig. 3). Surface charge screening also does not explain the differences in the Zn2+ effect between the subtypes. Mg2+ and Zn2+ are identically charged and contribute to surface charge screening equally. Mg2+ preferentially blocks Ba2+ currents more than Ca2+ currents through α1G (43), whereas we showed that Zn2+ block of α1G was comparable in Ba2+ and Ca2+. Thus, neither hydration energy nor surface charge screening alone appears to explain the influence of ion species–dependent modulation of channel conductance and Zn2+ block.

The selectivity filter of VDCCs alone does not explain the differences in the ion-dependent Zn2+ block effect. The molecular determinants of ion selectivity in VDCCs resides on the selectivity filter loci encoding EEEE/EEDD motif; however, α1E encoded with EEEE locus seen in HVA channels, as an exception, exhibits a higher permeability to Ca2+ than to Ba2+, similar to LVA channels (38). Thus, additional residues other than the EEEE/EEDD motif are involved in ion selectivity of the pore. Previous studies in K+ (44) and Na+ (45) channels demonstrated that metal binding sites are dependent on the unique residues near the selectivity filter. Mutation of these residues substantially affected Zn2+ block. These findings bring into question whether Ca2+ channels may also contain specific extrapore sites regulating Zn2+ block.

Zn2+ blockade is regulated by the putative EF-hand extrapore region of N-type calcium channel

We previously reported that ion permeation in N-type channels is modulated by a putative EF-hand motif located in domain III near the EEEE locus (26). It contains a central glycine residue flanked by three acidic residues, reminiscent of the classical EF-hands of Ca2+ binding proteins (46,47). Disruption of this motif reduces the ability of the channel to distinguish between Ba2+ and Ca2+ ions without affecting the pore function. To determine whether this putative EF-hand motif is also involved in regulating Zn2+ inhibition of these channels, we compared the IC50 values of Zn2+ obtained for wild-type N-type calcium channels to those obtained for two EF-hand mutant channels: a triple mutant in which all three negative charge residues in the putative EF-hand structure are replaced with positively charged residues (E1321K, D1323R, E1332R), and the G1326P mutant in which the central glycine is replaced with a proline. The latter mutant alters the ion permeability of the N-type channel by disrupting the EF-hand structure without changing the net local surface charge (26). We previously showed that these mutant channels exhibit biophysical properties similar to those of the wild-type channel (26). Representative recordings shown in Fig. 4 A reveal that Zn2+ inhibited the current activity of each mutant for both carriers, Ba2+ and Ca2+, in a dose-dependent manner. The mean Zn2+ IC50 values were comparable between the wild-type (98.0 ± 17.9 μM, n = 12) and mutant channels in 20 mM Ba2+ (triple, 79.2 ± 13.9 μM, n = 8; G1326P, 60.0 ± 5.5 μM, n = 11); however, the mean Zn2+ IC50 was significantly (p < 0.05) reduced in the mutant channels (wt, 1210 ± 80 μM, n = 8; triple, 290 ± 50 μM, n = 8; G1326P, 350 ± 70 μM, n = 11) when Ca2+ was used as the charge carrier (Fig. 4 B). Both the neutralization of three negatively charged residues and the replacement of central glycine with proline in the putative EF-hand region caused a similar effect on the Zn2+ sensitivity in Ca2+ condition, indicating that the effect is most likely independent of local surface charge screening. These results suggest that the putative EF-hand motif is involved in regulating Zn2+ block effect.

FIGURE 4.

FIGURE 4

Mutations of the putative EF-hand in domain III of the CaV2.2 (α1B) N-type channel affected Zn2+ sensitivity. (A) Representative current traces from two mutants of CaV2.2 (α1B) N-type calcium channels (trimutant, and G1326P) in the absence or presence of various concentrations of Zn2+, as indicated, in either 20 mM Ba2+ or 20 mM Ca2+. Trimutant: E1321K, D1323R, E1332R. (B) Summary of IC50 values of Zn2+ on the wild-type and mutant CaV2.2 (α1B) calcium channels. The data are presented as mean ± SE. (*) Statistical significance (p < 0.05) between the wild-type and mutant channels under the same recording conditions. The number in the parentheses indicates the number of cells used for the recordings.

The putative EF-hand in the domain III H5 loop of the N-type calcium channel α1B subunit forms a helix-coil-helix structure (48). A molecular structure model suggests that it is capable of interacting with Ca2+, although its sequence exhibits some variation compared to the classical EF-hand. The putative EF-hand was previously shown to be involved in Ba2+ and Ca2+ permeability of N-type channels (26) as well as ω-conotoxin GVIA block (49). Eight of the HVA Ca2+ channel α1 subunits appear to have this putative EF-hand motif (26), but none of the LVA Ca2+ channels, Na+ channels, and K+ channels have it in the homologous region (Fig. 5). Because mutation of the EF-hand motif of N-type channels increased Zn2+ sensitivity in Ca2+ toward that seen in α1G T-type channels, it may be involved in distinguishing between Zn2+ block of the HVA and LVA channels. In the typical EF-hand motif, the −Z position glutamic acid residue contributes two carboxylate oxygen atoms to coordinate a Ca2+, and the central glycine is located at sharp bend position critical for Ca2+ coordination (46,47). Shown in Fig. 5, both −Z acidic residue (CaV2.1, E1376; CaV2.2, E1332; CaV2.3, E1289; CaV1.2, E1081) and the central glycine (CaV2.1, E1370; CaV2.2, E1326; CaV2.3, E1283; CaV1.2, E1072) are conserved among four HVA Ca2+ channels; however, the remaining residues vary among the channels. In comparison to the CaV2 α1 subunits, CaV1.2/α1C contains three additional residues, K1078, D1079, and G1080, between the central glycine (G1072) and the −Z glutamic acid (E1081), which may weaken the effectiveness of the putative EF-hand motif in coordinating Ca2+. This could explain our finding that Zn2+ sensitivity of CaV1.2/α1C is higher than that of the CaV2 channels. Taken together, the presence and composition of the putative EF-hand motif may be partly responsible for the differences in Zn2+ block between HVA channels and LVA channels as well as between the HVA channels.

FIGURE 5.

FIGURE 5

Protein sequence alignments showing the extended pore regions of representative voltage-dependent calcium channels, sodium channels, and potassium channel. The box and asterisks indicate the identified selectivity filter residues. The solid line indicates the putative EF-hand motif region; solid circle, the conserved acidic residues and the central glycine; open circle, critical sites affecting Zn sensitivity of the channels. b, brain; s, skeletal muscle; and h, heart.

The differences in Zn2+ block effects on LVA channels do not appear to be explained by the putative EF-hand motif because it is absent in their α1 subunit. In this study, we found that under Ba2+ conditions α1H was ∼10-fold more sensitive to Zn2+ than α1G and α1I. In comparison, under Ca2+ conditions, α1H was ∼100- to 200-fold more sensitive to Zn2+ than α1G and α1I (18,28). These charge carrier-dependent differences in Zn2+ sensitivities indicate that block is modulated by residues outside the EEDD selectivity filter locus. Similarly, α1H has been shown to have higher sensitivities to Cu2+ and Ni2+ than the other T-type channels (18) through a mechanism that may involve noncharged residues. Cysteine or histidine pairs are known to be able to coordinate divalent ions and have been shown to affect ion selectivity when found in the pore regions of Na+ (45) and K+ channels (44). A recent report showed that the point mutation H191Q in the S3-S4 loop of domain I reduced Ni+ block of α1H (50), indicating that H191 is critical for Ni+ block. Whether the high-sensitivity block of α1H by Zn2+ also resides with this residue remains to be tested, but the study lends further support to the notion that extrapore residues can affect ion block of VDCCs.

In conclusion, we have systematically tested and compared the Zn2+ sensitivities of HVA and LVA calcium channels in a transient expression system. We show that Zn2+ block of VDCCs is dependent on the pore-forming α1 subunit, but variation in the Zn2+ sensitivities of the channel subtypes is independent of EEEE/EEDD locus and is differentially regulated by permeant ions. Zn2+ block of the CaV2.2/α1B N-type channel is modulated via a putative EF-hand motif outside of the EEEE locus. Although it is not clear how the EF-hand is involved in modulation of Zn2+ block, we envision a model in which Zn2+ binds at the pore of the channel, and other regions, such as the EF-hand, modulate the interaction between Zn2+ and the pore. Our study supports the notion that multiple mechanisms are involved in the channel subtype-dependent Zn2+ block. Further studies are required to identify other molecular determinants underlying the differences in the block effect of Zn2+ on the various calcium channel subtypes and to elucidate the mechanisms involving these determinants.

Acknowledgments

We thank Dr. Terry Snutch and Dr. Gerald Zamponi for the calcium channel constructs. We thank Clinton Doaring for his technical assistance.

This work was supported by an operating grant to Z.P.F. from the Heart and Stroke Foundation of Ontario (grant No. NA 5302), startup funds from the Dept. of Physiology, the Faculty of Medicine, and the University of Toronto, and equipment grants from the Canada Foundation for Innovation and the Ontario Innovation Trust. Z.P.F. holds a New Investigator Award from the Canadian Institutes of Health Research and a Premier's Research Excellence Award of Ontario.

Editor: Eduardo Perozo.

References

  • 1.Catterall, W. A. 2000. Structure and regulation of voltage-gated Ca2+ channels. Annu. Rev. Cell Dev. Biol. 16:521–555. [DOI] [PubMed] [Google Scholar]
  • 2.Reuter, H., S. Kokubun, and B. Prod'hom. 1986. Properties and modulation of cardiac calcium channels. J. Exp. Biol. 124:191–201. [DOI] [PubMed] [Google Scholar]
  • 3.Takeda, A., A. Minami, Y. Seki, and N. Oku. 2004. Differential effects of zinc on glutamatergic and GABAergic neurotransmitter systems in the hippocampus. J. Neurosci. Res. 75:225–229. [DOI] [PubMed] [Google Scholar]
  • 4.Assaf, S. Y., and S. H. Chung. 1984. Release of endogenous Zn2+ from brain tissue during activity. Nature. 308:734–736. [DOI] [PubMed] [Google Scholar]
  • 5.Frederickson, C. J., L. J. Giblin III, B. Rengarajan, R. Masalha, C. J. Frederickson, Y. Zeng, E. V. Lopez, J. Y. Koh, U. Chorin, L. Besser, M. Hershfinkel, Y. Li, R. B. Thompson, and A. Krezel. 2006. Synaptic release of zinc from brain slices: factors governing release, imaging, and accurate calculation of concentration. J. Neurosci. Methods. 154:19–29. [DOI] [PubMed] [Google Scholar]
  • 6.Smart, T. G., X. Xie, and B. J. Krishek. 1994. Modulation of inhibitory and excitatory amino acid receptor ion channels by zinc. Prog. Neurobiol. 42:393–441. [DOI] [PubMed] [Google Scholar]
  • 7.Lin, D. D., A. S. Cohen, and D. A. Coulter. 2001. Zinc-induced augmentation of excitatory synaptic currents and glutamate receptor responses in hippocampal CA3 neurons. J. Neurophysiol. 85:1185–1196. [DOI] [PubMed] [Google Scholar]
  • 8.Koh, J. Y., S. W. Suh, B. J. Gwag, Y. Y. He, C. Y. Hsu, and D. W. Choi. 1996. The role of zinc in selective neuronal death after transient global cerebral ischemia. Science. 272:1013–1016. [DOI] [PubMed] [Google Scholar]
  • 9.Turan, B., H. Fliss, and M. Desilets. 1997. Oxidants increase intracellular free Zn2+ concentration in rabbit ventricular myocytes. Am. J. Physiol. 272:H2095–H2106. [DOI] [PubMed] [Google Scholar]
  • 10.Bagate, K., J. J. Meiring, M. E. Gerlofs-Nijland, F. R. Cassee, H. Wiegand, A. Osornio-Vargas, and P. J. Borm. 2006. Ambient particulate matter affects cardiac recovery in a Langendorff ischemia model. Inhal. Toxicol. 18:633–643. [DOI] [PubMed] [Google Scholar]
  • 11.Graff, D. W., W. E. Cascio, J. A. Brackhan, and R. B. Devlin. 2004. Metal particulate matter components affect gene expression and beat frequency of neonatal rat ventricular myocytes. Environ. Health Perspect. 112:792–798. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Busselberg, D., B. Platt, D. Michael, D. O. Carpenter, and H. L. Haas. 1994. Mammalian voltage-activated calcium channel currents are blocked by Pb2+, Zn2+, and Al3+. J. Neurophysiol. 71:1491–1497. [DOI] [PubMed] [Google Scholar]
  • 13.Easaw, J. C., B. S. Jassar, K. H. Harris, and J. H. Jhamandas. 1999. Zinc modulation of ionic currents in the horizontal limb of the diagonal band of Broca. Neuroscience. 94:785–795. [DOI] [PubMed] [Google Scholar]
  • 14.Noh, J. H., and J. M. Chung. 2003. Zinc reduces low-threshold Ca2+ currents of rat thalamic relay neurons. Neurosci. Res. 47:261–265. [DOI] [PubMed] [Google Scholar]
  • 15.Nam, S. C., and P. E. Hockberger. 1992. Divalent ions released from stainless steel hypodermic needles reduce neuronal calcium currents. Pflugers Arch. 420:106–108. [DOI] [PubMed] [Google Scholar]
  • 16.Muller, T. H., U. Misgeld, and D. Swandulla. 1992. Ionic currents in cultured rat hypothalamic neurones. J. Physiol. 450:341–362. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Malaiyandi, L. M., K. E. Dineley, and I. J. Reynolds. 2004. Divergent consequences arise from metallothionein overexpression in astrocytes: zinc buffering and oxidant-induced zinc release. Glia. 45:346–353. [DOI] [PubMed] [Google Scholar]
  • 18.Jeong, S. W., B. G. Park, J. Y. Park, J. W. Lee, and J. H. Lee. 2003. Divalent metals differentially block cloned T-type calcium channels. Neuroreport. 14:1537–1540. [DOI] [PubMed] [Google Scholar]
  • 19.Magistretti, J., L. Castelli, V. Taglietti, and F. Tanzi. 2003. Dual effect of Zn2+ on multiple types of voltage-dependent Ca2+ currents in rat palaeocortical neurons. Neuroscience. 117:249–264. [DOI] [PubMed] [Google Scholar]
  • 20.Beedle, A. M., J. Hamid, and G. W. Zamponi. 2002. Inhibition of transiently expressed low- and high-voltage-activated calcium channels by trivalent metal cations. J. Membr. Biol. 187:225–238. [DOI] [PubMed] [Google Scholar]
  • 21.Santi, C. M., F. S. Cayabyab, K. G. Sutton, J. E. McRory, J. Mezeyova, K. S. Hamming, D. Parker, A. Stea, and T. P. Snutch. 2002. Differential inhibition of T-type calcium channels by neuroleptics. J. Neurosci. 22:396–403. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Stea, A., T. W. Soong, and T. P. Snutch. 1995. Voltage-gated calcium channels. In Ligand- and Voltage-Gated Ion Channels. R.A.North, editor. CRC Press, Boca Raton, FL. 113–141.
  • 23.Birnbaumer, L., K. P. Campbell, W. A. Catterall, M. M. Harpold, F. Hofmann, W. A. Horne, Y. Mori, A. Schwartz, T. P. Snutch, T. Tanabe, et al. 1994. The naming of voltage-gated calcium channels. Neuron. 13:505–506. [DOI] [PubMed] [Google Scholar]
  • 24.Ertel, E. A., K. P. Campbell, M. M. Harpold, F. Hofmann, Y. Mori, E. Perez-Reyes, A. Schwartz, T. P. Snutch, T. Tanabe, L. Birnbaumer, R. W. Tsien, and W. A. Catterall. 2000. Nomenclature of voltage-gated calcium channels. Neuron. 25:533–535. [DOI] [PubMed] [Google Scholar]
  • 25.McGivern, J. G. 2006. Pharmacology and drug discovery for T-type calcium channels. CNS Neurol. Disord. Drug Targets. 5:587–603. [DOI] [PubMed] [Google Scholar]
  • 26.Feng, Z. P., J. Hamid, C. Doering, S. E. Jarvis, G. M. Bosey, E. Bourinet, T. P. Snutch, and G. W. Zamponi. 2001. Amino acid residues outside of the pore region contribute to N-type calcium channel permeation. J. Biol. Chem. 276:5726–5730. [DOI] [PubMed] [Google Scholar]
  • 27.Hui, K., P. Gardzinski, H. S. Sun, P. H. Backx, and Z. P. Feng. 2005. Permeable ions differentially affect gating kinetics and unitary conductance of L-type calcium channels. Biochem. Biophys. Res. Commun. 338:783–792. [DOI] [PubMed] [Google Scholar]
  • 28.Traboulsie, A., J. Chemin, M. Chevalier, J. F. Quignard, J. Nargeot, and P. Lory. 2007. Subunit-specific modulation of T-type calcium channels by zinc. J. Physiol. 578:159–171. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Cribbs, L. L., J. H. Lee, J. Yang, J. Satin, Y. Zhang, A. Daud, J. Barclay, M. P. Williamson, M. Fox, M. Rees, and E. Perez-Reyes. 1998. Cloning and characterization of alpha1H from human heart, a member of the T-type Ca2+ channel gene family. Circ. Res. 83:103–109. [DOI] [PubMed] [Google Scholar]
  • 30.Perez-Reyes, E., L. L. Cribbs, A. Daud, A. E. Lacerda, J. Barclay, M. P. Williamson, M. Fox, M. Rees, and J. H. Lee. 1998. Molecular characterization of a neuronal low-voltage-activated T-type calcium channel. Nature. 391:896–900. [DOI] [PubMed] [Google Scholar]
  • 31.Lee, J. H., A. N. Daud, L. L. Cribbs, A. E. Lacerda, A. Pereverzev, U. Klockner, T. Schneider, and E. Perez-Reyes. 1999. Cloning and expression of a novel member of the low voltage-activated T-type calcium channel family. J. Neurosci. 19:1912–1921. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Hess, P., and R. W. Tsien. 1984. Mechanism of ion permeation through calcium channels. Nature. 309:453–456. [DOI] [PubMed] [Google Scholar]
  • 33.Heinemann, S. H., H. Terlau, W. Stuhmer, K. Imoto, and S. Numa. 1992. Calcium channel characteristics conferred on the sodium channel by single mutations. Nature. 356:441–443. [DOI] [PubMed] [Google Scholar]
  • 34.McDonald, T. F., S. Pelzer, W. Trautwein, and D. J. Pelzer. 1994. Regulation and modulation of calcium channels in cardiac, skeletal, and smooth muscle cells. Physiol. Rev. 74:365–507. [DOI] [PubMed] [Google Scholar]
  • 35.Yue, D. T., and E. Marban. 1990. Permeation in the dihydropyridine-sensitive calcium channel. Multi-ion occupancy but no anomalous mole-fraction effect between Ba2+ and Ca2+. J. Gen. Physiol. 95:911–939. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Nilius, B., P. Hess, J. B. Lansman, and R. W. Tsien. 1985. A novel type of cardiac calcium channel in ventricular cells. Nature. 316:443–446. [DOI] [PubMed] [Google Scholar]
  • 37.Carbone, E., and H. D. Lux. 1984. A low voltage-activated, fully inactivating Ca channel in vertebrate sensory neurones. Nature. 310:501–502. [DOI] [PubMed] [Google Scholar]
  • 38.Bourinet, E., G. W. Zamponi, A. Stea, T. W. Soong, B. A. Lewis, L. P. Jones, D. T. Yue, and T. P. Snutch. 1996. The alpha 1E calcium channel exhibits permeation properties similar to low-voltage-activated calcium channels. J. Neurosci. 16:4983–4993. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Hess, P., J. B. Lansman, and R. W. Tsien. 1986. Calcium channel selectivity for divalent and monovalent cations. Voltage and concentration dependence of single channel current in ventricular heart cells. J. Gen. Physiol. 88:293–319. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Busselberg, D., D. Michael, M. L. Evans, D. O. Carpenter, and H. L. Haas. 1992. Zinc (Zn2+) blocks voltage gated calcium channels in cultured rat dorsal root ganglion cells. Brain Res. 593:77–81. [DOI] [PubMed] [Google Scholar]
  • 41.Marcus, Y. 1994. A simple empirical-model describing the thermodynamics of hydration of ions of widely varying charges, sizes, and shapes. Biophys. Chem. 51:111–127. [Google Scholar]
  • 42.Hefter, G., Y. Marcus, and W. E. Waghorne. 2002. Enthalpies and entropies of transfer of electrolytes and ions from water to mixed aqueous organic solvents. Chem. Rev. 102:2773–2836. [DOI] [PubMed] [Google Scholar]
  • 43.Serrano, J. R., S. R. Dashti, E. Perez-Reyes, and S. W. Jones. 2000. Mg2+ block unmasks Ca2+/Ba2+ selectivity of alpha1G T-type calcium channels. Biophys. J. 79:3052–3062. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Krovetz, H. S., H. M. VanDongen, and A. M. VanDongen. 1997. Atomic distance estimates from disulfides and high-affinity metal-binding sites in a K+ channel pore. Biophys. J. 72:117–126. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Backx, P. H., D. T. Yue, J. H. Lawrence, E. Marban, and G. F. Tomaselli. 1992. Molecular localization of an ion-binding site within the pore of mammalian sodium channels. Science. 257:248–251. [DOI] [PubMed] [Google Scholar]
  • 46.Tufty, R. M., and R. H. Kretsinger. 1975. Troponin and parvalbumin calcium binding regions predicted in myosin light chain and T4 lysozyme. Science. 187:167–169. [DOI] [PubMed] [Google Scholar]
  • 47.da Silva, A. C., and F. C. Reinach. 1991. Calcium binding induces conformational changes in muscle regulatory proteins. Trends Biochem. Sci. 16:53–57. [DOI] [PubMed] [Google Scholar]
  • 48.Doughty, S. W., F. E. Blaney, B. S. Orlek, and W. G. Richards. 1998. A molecular mechanism for toxin block in N-type calcium channels. Protein Eng. 11:95–99. [DOI] [PubMed] [Google Scholar]
  • 49.Feng, Z. P., J. Hamid, C. Doering, G. M. Bosey, T. P. Snutch, and G. W. Zamponi. 2001. Residue Gly1326 of the N-type calcium channel alpha 1B subunit controls reversibility of omega-conotoxin GVIA and MVIIA block. J. Biol. Chem. 276:15728–15735. [DOI] [PubMed] [Google Scholar]
  • 50.Kang, H. W., J. Y. Park, S. W. Jeong, J. A. Kim, H. J. Moon, E. Perez-Reyes, and J. H. Lee. 2006. A molecular determinant of nickel inhibition in CaV3.2 T-type calcium channels. J. Biol. Chem. 281:4823–4830. [DOI] [PubMed] [Google Scholar]

Articles from Biophysical Journal are provided here courtesy of The Biophysical Society

RESOURCES