Main

The electrophysiological and pharmacological diversity of native calcium (Ca2+) channels (L, N, P, Q and T types) is well documented1. These subtypes have different functions; low-voltage-activated T-type channels shape action potentials and generate firing patterns2, whereas L-type channels regulate Ca2+-dependent genes and enzymes3,4, and N-type and P/Q-type channels contribute to neurotransmitter release5,6,7,8.

High-threshold neuronal Ca2+ channels are heterotrimeric complexes composed of a pore-forming α1 subunit associated with β and α2δ subunits. Each of the cloned α1 subunits has distinct functional characteristics that are modulated by co-expression of any of four β subunits9. Channels formed from α1C and α1D subunits have the properties of neuronal dihydropyridine-sensitive L-type channels, and α1B channels encode ω-CgTx GVIA-sensitive N-type channels9. The α1E channel shares some properties with both low-threshold channels (high sensitivity to nickel block, permeation Ca2+ > Ba2+) and high-threshold Ca2+ channels (activation at more positive potentials, single-channel conductance > 10 pS)10,11,12, although the native counterpart of this subunit remains unclear. T-type low-threshold channels are encoded by at least three different α1 subunits (α1G, α1H, α1I) and seem not to require β and α2δ subunits for functional expression13,14.

P-type Ca2+ channels, originally described in cerebellar Purkinje cells15, are widely distributed16,17 and mediate neurotransmitter release in the central and peripheral nervous systems5,6,7,8. Q-type Ca2+ channels, first described in cerebellar granule cells18,19, also mediate neurotransmitter release at some synapses20. Native P- and Q-type channels differ in sensitivity to ω-agatoxin IVA (ω-Aga IVA; Kd ~2 nM for P-type versus > 100 nM for Q-type) and in their inactivation kinetics. (P-type currents show a non-inactivating waveform during prolonged membrane depolarization, whereas Q-type currents show a pronounced inactivation.) Some native Ca2+ channels have properties related to but distinct from either P- or Q-type channels, suggesting a closely related family of channel types21,22.

The α1A subunit23,24 is highly expressed in both Purkinje and cerebellar granule cells and shares several properties with both native P- and Q-type Ca2+ currents25,26. Mutations in the α1A subunit result in several neuropathological conditions, including familial hemiplegic migraine, cerebellar ataxia and epilepsy27,28,29. The properties of α1A do not exactly match those of either P- or Q-type Ca2+ currents, and there has also been no account for their distinct sensitivities to ω-Aga IVA. Here we report that alternative splicing at distinct sites within the α1A subunit gene generate multiple types of P- and Q-type conductances that exhibit distinct gating, pharmacology and modulatory characteristics. We also suggest there is little reason that α1A channels should be selectively subdivided into distinct P- versus Q- subtypes.

Results

α1A isoforms are generated by alternative splicing

Screening of a rat brain cDNA library identified a calcium channel α1A subunit (called α1A-b) that differed from the primary sequence of the original rat brain α1A-a ( ref. 24) at three sites ( Fig. 1a ). The domain I–II linker in α1A-b contained a single valine insertion (Val421) located 18 residues carboxyl to the β subunit binding site30. In domain IV, α1A-b contained an insertion of two residues, Asp and Pro (N1605-P1606), in the extracellular linker separating transmembrane segments S3 and S4. Finally, α1A-b contained ten substituted residues in a stretch of thirty amino acids in the carboxyl tail adjacent to domain IV S6. Of the ten substitutions, six are located in a region highly similar to the divalent ion binding domain (EF hand) of Ca2+-binding proteins31.

Figure 1: Alternative splicing generates multiple α1A isoforms.
figure 1

(a) Differences between α1A-a and α1A-b subunits (single-letter code). (b) Comparison of the α1A genomic and cDNA sequences in the domain I–II linker identifies possible donor and acceptor sites. Exons are indicated by filled boxes and intron sequences by lines. (c) Comparison of the α1A genomic and cDNA sequences in the domain IV S3–S4 region shows that a 162-bp intron separates two exons.

To determine the molecular nature of these differences, we sequenced clones from a rat genomic library and also analyzed rat DNA by PCR. Clones corresponding to the region flanking the α1A I–II linker had an Asp preceding the Val421 insertion ( Fig. 1b ). A 1.8-kb intron separated the Asp residue from downstream Gly422 found in α1A-a transcripts. The flanking genomic and cDNA sequences contained a common 5´ splice donor site and three possible 3´ acceptor sites: one that would result in either the inclusion or exclusion of Val421, another that would include Gly422, and a third that could produce an α1A variant lacking both the Gly and Val421 residues ( Fig. 1b ). To test this third possibility, we examined five additional α1A cDNAs and found that one of five lacked both the Val421 and Gly422 residues (called α1A-c). To confirm that the α1A-c variant was expressed in vivo, we examined total cerebellar and hippocampal RNAs by RT-PCR followed by DNA sequencing. Of 81 PCR products, 16% were the α1A-c variant, 79% the domain I–II-linker α1A-a isoform, and 5% the α1A-b isoform (data not shown).

PCR amplification of rat genomic DNA across the domain IV S3–S4 region identified a 162-bp intron that when absent resulted in α1A channels lacking N1605-P16061A-a variant), whereas inclusion of N1605-P1606 (the α1Ab isoform) resulted from use of an alternative 5´ splice donor site ( Fig. 1c ). The genomic nature of the differences between α1A-a and α1Ab in the EF hand region were not determined, although the 10 amino-acid substitutions are contained in a contiguous stretch of 90 bp with only 53% sequence identity, suggesting a mutually exclusive alternative splicing mechanism.

Differential expression of α1A isoforms in the rat CNS

The spatial expression patterns of α1A-a and α1Ab in the rat CNS were examined by RT-PCR of the domain I–II linker region and by in situ hybridization using antisense oligonucleotide probes against the distinct EF hand sequences. Stringent PCR conditions allowed the selective amplification of α1A-b variants containing the 3-bp Val421 insertion in the domain I–II linker ( Fig. 2a ). The expected α1A-b 208-bp fragment was detected in all rat brain regions examined, indicating that α1A transcripts containing Val421 are widely expressed. Furthermore, transcripts containing Val421 were also detected at moderate levels in kidney RNA ( Fig. 2a ). Attempts to amplify the 208-bp fragment from rat genomic DNA samples were unsuccessful, consistent with the presence of the intronic sequences in the α1A I–II linker (see Fig. 1b ).

Figure 2: Both α1A-a and α1A-b variants are expressed in the rat nervous system.
figure 2

(a) RT-PCR demonstrating the expression of α1A-b transcripts. Blotting and hybridization of the PCR products with an internal oligonucleotide showed the specific amplification of an α1A-b 208-bp fragment from the α1A-b template cDNA control but not the α1A-a cDNA. (b–e) In-situ hybridization to adult rat brain sections using radiolabeled oligonucleotide probes specific to the α1A-a and α1A-b EF hand regions. Sections show α1A-b (b) and α1A-a (c) in cerebellum and α1A-b (d) and α1A-a (e) in hippocampal CA1-CA3 pyramidal (p) and dentate granule cells (g).

In situ hybridization with the EF hand probes revealed differential expression of α1A-a and α1A-b transcripts. Most brain regions prominently expressed α1A-b. Expression of α1A-a was strong in the cerebellar cortex, with lower levels in the cortex, hippocampus and olfactory tubercule (not shown). Both α1A-a and α1A-b were detected in cerebellar granule cells, with α1A-a predominantly in Purkinje cells ( Fig. 2b and c ). In contrast, α1A-b was highly expressed in CA1-CA3 pyramidal and dentate granule cells in the hippocampus, where α1A-a expression was significantly lower ( Fig. 2d and e ).

Transcripts encoding α1A variants with N1605-P1606 contain an additional Tfi I restriction enzyme site. PCR amplication of rat cerebellar and hippocampal α1A RNAs and digestion with Tfi I showed that both regions expressed both α1A channel variants. However, in cerebellar RNA, most isoforms lacked N1605-P1606, whereas in hippocampus both forms were represented approximately equally (not shown).

α1A splicing affects electrophysiological properties

The electrophysiological properties of α1A-a and α1A-b channels were compared by transient expression in Xenopus oocytes (co-expressed with rat brain α2δ and β4 subunits). The α1A-b isoform had three significant functional differences from α1A-a ( Fig. 3 ). First, α1A-b inactivation rate was dramatically slowed, with 84 ± 3% (n = 12) of the whole-cell current remaining after a 400-ms test pulse compared to 39 ± 8% for α1A-a (n = 8; Fig. 3a ). After a 16-s test pulse, approximately 50% of the α1A-b current remained, whereas α1A-a currents were completely inactivated ( Fig. 3b ). Second, α1A-b showed a significant (p < 0.05) positive shift in the current–voltage relationship by ~6 mV (α1A-a V0.5 = –4.1 ± 0.5, n = 18; α1A-b V0.5 = 2.1 ± 0.7, n = 20; Fig. 3a ). Third, the voltage dependence of inactivation of α1A-b was shifted more positively by ~20 mV. Furthermore, even after the holding potential was clamped to +50 mV for 15 s, ~60% of the whole-cell current was not inactivated ( Fig. 3c ).

Figure 3: α1A-a and α1A-b have distinct functional properties.
figure 3

(a) Waveforms and current–voltage relationships for α1A-a and α1A-b. Waveforms of peak normalized current from a holding potential of –100 mV show that α1A-b barely inactivates over 400 ms. The mean normalized I–V curves for α1A-a (n = 18) and α1A-b ( n = 20) show a positive shift for α1A-b. (b) Normalized waveforms for α1A-a and α1A-b channels co-expressed with three different Ca2+ channel β subunits. Cells were depolarized from a holding potential of –100 mV to the peak I–V potential, and the currents remaining after 2 s or 16 s were averaged and presented as histograms (mean, n = 8–17 cells for each combination). (c) Steady-state inactivation curves for α1A-a and α1A-b co-expressed with β4 and α2. Cells were held at various potentials for 16 s before a test pulse to +10 mV. The half-inactivation potentials obtained from the fit were –17.2 ± 0.7 mV (n = 5) for α1A-a and –1.6 ± 0.8 mV (n = 10) for α1A-b.

Because co-expression of Ca2+ channel β subunits both alters channel kinetics and shifts the current–voltage relationship9, we asked whether these functional differences were specific for neuronal β subunit types. The inactivation rate of α1A-b was significantly slower than α1A-a for all β subunits tested ( Fig. 3b ), regardless of whether the β subunit normally increases (β1b) or decreases (β2a) inactivation rates. Furthermore, regardless of the co-expressed β subunit, the current–voltage relationship of α1A-b was significantly (p < 0.05) more positive when compared to α1A-a (not shown). The amino-acid differences in α1A-b were not associated with significant changes in other channel properties examined, including ion permeation (not shown), the rate of activation (not shown) and the slope of the current–voltage relationship (α1A-a, k = –3.5 ± 0.2, n = 18; α1A-b, k = –3.7 ± 0.5, n = 20).

Val421 and N1605-P1606 differentially affect properties

To identify the regions of α1A-b responsible for its distinct gating characteristics, we constructed chimeras between α1A-a and α1A-b ( Fig. 4a ). The α1A-a (+V) chimera was identical to α1A-a except for the Val421 insertion in the domain I–II linker, and the α1A-b (–V) chimera differed from α1A-a only by insertion of N1605-P1606 and substitution of the α1A-b EF hand region. The α1A-b (–V) chimera showed fast inactivation kinetics similar to those of α1A-a (42 ± 3% remaining after 400 ms, n = 9) but retained a current–voltage relationship positively shifted by ~6 mV (V0.5 = 2.7 ± 0.5, n = 19; Fig. 3a ). In contrast, the α1A-a (+V) chimera had slow inactivation kinetics similar to that of α1A-b (79 ± 3% remaining after 400 ms, n = 13) and a relatively negative current–voltage relationship similar to α1A-a (V0.5 = –5.0 ± 0.3, n = 21; Fig. 3a ).

Figure 4: Splicing of the domain I–II linker and domain IV S3–S4 loop make distinct contributions to channel kinetics and gating.
figure 4

(a) Schematic representation of chimeric α1A Ca2+ channels and their inactivation and activation properties in the presence of β4 and α2 (n = 9–21). Inactivation is measured as the percentage of whole-cell peak current remaining at the end of a 400-ms test pulse and activation as half-activation potentials obtained from fits to whole-cell current–voltage relationships. (b) Single-channel records obtained from cell-attached patch recordings from oocytes expressing α1A-a with or without Val421 (+ β4 and α2). Currents were elicited from a holding potential of –100 mV to a test potential of 0 mV. Bursts of activity separated by brief quiescent periods were observed in α1A-a (+V). (c) Effect of Val421 on the mean open time distributions within individual bursts. Without Val421, the open time distribution was fitted with two exponentials with time constants of 0.34 ms (44%) and 1.33 ms (56%). With Val421, time constants were 0.90 ms (35%), 3.26 ms (51%) and 15.74 ms (14%).

We examined two additional chimeras: α1A-a (+NP), which was identical to α1A-a except for the insertion of N1605-P1606, and α1A-b (–NP), which was identical to α1A-b except that it lacked N1605-P1606 ( Fig. 4a ). Inactivation properties of α1A-a (+NP) were identical to those of α1A-a, although V0.5 was shifted similarly to α1A-b1A-a (+NP) = 3.3 ± 0.4, n = 11). In contrast, the α1A-b (–NP) chimera had negatively shifted V0.5 but slow inactivation kinetics typical of the wild-type α1A-b1A-b (–NP) = –3.9 ± 0.4, n = 12). Another chimera containing the α1A-c variation in the domain I–II linker (α1A-a missing both Val421 and Gly422) showed similar waveform and inactivation properties to α1A-a and was not examined further in the present study (not shown). We could not detect any distinct contributions of the spliced EF hand regions to any of the Ca2+ channel properties examined in this study.

Presence or absence of Val421 affects inactivation

Cell-attached, single-channel recordings were made in oocytes expressing either the α1A-a or the α1A-a (+V) chimera. At the single-channel level, the presence of Val421 caused the appearance of multiple bursts of channel activity ( Fig. 4b ), which continued throughout test pulses as long as 24 s (not shown). Ensemble averages of sweeps with unitary activity showed a rapidly inactivating α1A-a current, whereas α1A-a (+V) averaged currents were sustained ( Fig. 4b ). Varying the holding potential from –100 to –20 mV completely inactivated α1A-a channels but did not affect α1A-a (+V) open probability, consistent with the lack of complete steady-state inactivation at the whole-cell level (not shown).

Comparison of the open-time distributions for α1A-a and α1A-a (+V) channels ( Fig. 4c ) shows that α1A-a openings are characterized by two apparent brief open states (0.34 and 1.33 ms), whereas the α1A-a (+V) distribution had three markedly longer open states (0.90, 3.26 and 15.74 ms) . Thus, Val421 contributes to two separate effects: destabilization of the inactivated state, as indicated by the appearance of multiple bursts, and stabilization of the open state(s), as indicated by the increase in mean open time and appearance of an additional time constant. There was no significant change in the single-channel slope conductance of α1A channels with or without Val421 (not shown; α1A-a, 16.5 ± 0.4 pS, α1A-a(+V), 17.2 ± 0.7 pS; p < 0.05, n = 4).

N1605-P1606 reduces ω-agatoxin IVA sensitivity

The major distinguishing characteristic between native P- and Q-type calcium currents is their differential sensitivity to the spider peptide toxin ω-Aga IVA8,16,17,18,19,19. In Xenopus oocytes, α1A-a and α1A-b channels were blocked approximately equally by 200 nM ω-Aga IVA (~25% block) and by 5 μM ω-CTx-MVIIC (~80% block; Fig. 5a ). In contrast to native P-type channels, ω-Aga IVA block was not reversed by application of depolarizing prepulses (data not shown). However, in transfected human embryonic kidney (HEK 293) cells, 100 to 300 nM ω-Aga IVA completely blocked the wild-type α1A-a ( Fig. 5b and c ). As for native P-type channels, block was relieved by a series of strong depolarizations ( Fig. 5b ). The sensitivity of α1A-a to ω-Aga IVA was markedly reduced in channels containing N1605-P1606 in domain IV S3–S4 ( Fig. 5c ). The N1605-P1606 insertion lowered the blocking rate constant, kon, 7-fold (α1A-a = 6.475 × 10–3 μM–1s–1; α1A-a (+NP) = 0.9198 × 10–3 μM–1s–1) and increased koff 2-fold (α1A-a = 1.47 × 10–3 s–1; α1A-a (+NP) = 3.045 × 10–3 s–1), resulting in an 11-fold decrease in affinity for ω-Aga IVA ( Fig. 5d and e ; Kd for α1A-a = 14.8 nM; Kd for α1A-a (+NP) = 167 nM). In contrast to ω-Aga IVA, the snail peptide toxin ω-CTx-MVIIC blocked both α1A-a and α1A-a (+NP) with a similar time course and potency ( Fig. 5f ). These pharmacological profiles in HEK cells suggest that native P-type channels lack N1605-P1606, whereas Q-type channels contain these two residues.

Figure 5: Sensitivities of α1A-a and α1A-a (+NP) to ω-Aga-IVA and ω-CTx-MVIIC.
figure 5

Whole-cell current records obtained from α1A-a and α1A-b (expressed transiently in Xenopus oocytes) after application of 200 nM ω-Aga-IVA or 6 μM ω-CTx-MVIIC. Mean ± s.e.; n = 5–8. (b–f) Channels expressed transiently in HEK cells. (b) Time course of development of ω-Aga IVA (300 nM) block of α1A-a and reversal of toxin action with repetitive depolarizations (10 × 10 ms steps to +120 mV at 1 Hz). Test pulses to 0 mV were elicited from a holding potential of –100 mV every 15 seconds. (c) Whole-cell current records obtained from α1A-a and α1A-a (+NP) with or without 100 nM ω-Aga IVA. (d) Average normalized time course comparing the rate of development of ω-Aga IVA (300 nM) block of α1A-a and α1A-a (+NP). Inset, on-rate of block for α1A-a and α1A-a (+NP) varies linearly with ω-Aga IVA concentration. Mean data (± s.e.; n = 3–5) were fit according to the equation 1/τon =kon [ω-Aga IVA] + koff. The Kd values were 14.8 nM for α1A-a and 167 nM for α1A-a (+NP). (e) Dose dependence of the effects of ω-Aga IVA for α1A-a and α1A-a (+NP). The solid lines reflect a fit with the Hill equation with the Hill coefficient arbitrarily held at 1.0. The IC50 values obtained were 16.3 nM for α1A-a and 146 nM for α1A-a (+NP). (f) Time course of development of ω-CTx-MVIIC (6 μM) block of α1A-a and α1A-a (+NP). Inset, blocking rate constants obtained for α1A-a and α1A-a (+NP). Test pulses to +10 mV were elicited from a holding potential of –100 mV every 15 seconds.

G proteins and PKC modulate α1A splice variants

The activation of certain neurotransmitter receptors inhibits P/Q-type Ca2+ channels through the binding of G-protein βγ subunits directly to the α1A domain I–II linker32. G-protein modulation of Ca2+ channels characteristically involves decreased whole-cell currents, slowed activation and inactivation kinetics, and relief from inhibition by application of positive prepulses (called facilitation)33. Because Val421 is located in one of two Gβγ binding sites of the I–II linker and may directly alter Gβγ binding affinity, we tested whether this residue affects the G-protein-dependent modulation of α1A channels. The μ-opioid receptor agonist DAMGO inhibited both α1A-a and α1A-b currents in cells co-expressing the μ-opioid receptor ( Fig. 6 ). However, DAMGO-inhibited α1A-b activation kinetics were significantly more slowed and the overall opioid-induced inhibition more pronounced compared with α1A-a ( Fig. 6a ). Times to peak indicated that Val421 is responsible for the slowed kinetics of α1A-b after G-protein activation ( Fig. 6a ). In the presence of DAMGO, the two splice variants also showed distinct magnitudes of voltage-dependent facilitation as determined by applying a 50-ms prepulse to +150 mV before a test pulse from –100 mV to + 10 mV. When measured 20 ms after the test pulse, the α1A-b variant showed twofold more facilitation (α1A-b, 22 ± 3%, n = 15; α1A-a, 10 ± 2%, n = 14; data not shown). Furthermore, varying the time between the prepulse and the test pulse showed that the α1A-a-facilitated currents were re-inhibited more rapidly compared to α1A-b ( Fig. 6b ; τ reinhibition α1A-a = 25 ± 2 ms, n = 15; α1A-a(+V), 39 ± 3, n = 14). The difference in re-inhibition kinetics following DAMGO-induced G-protein inhibition is due to Val421 in the domain I–II linker ( Fig. 6b ).

Figure 6: Differential modulation of α1A splice variants by G proteins and PKC.
figure 6

(a, b) Modulation of α1A-a and α1A-b by G proteins through the activation of a co-expressed μ-opioid receptor. (a) Currents elicited from a holding potential of –100 mV to a test potential of +10 mV with or without 1 μM DAMGO. Time-to-peak measurement shows that Val421 is responsible for the slower kinetics of α1A-b following μ-opioid receptor activation. (b) Time course of reinhibition of the channels by G-proteins after temporary relief with a strong depolarizing prepulse to +150 mV. The presence of Val421 slows the reinhibition kinetics, suggesting a lower affinity of the channel for Gβγ. (c, d) Activation of PKC-dependent phosphorylation by application of 100 nM PMA. PKC-dependent upregulation of α1A-a and α1A-b follows a similar time course (c), but the presence of Val421 causes a twofold increase in the upregulation (c, d). In (a), α1A was co-expressed with β4 and α2 in Xenopus oocytes. In (b–d), α1A was co-expressed with β4 and α2 in HEK 293 cells.

The stimulation of protein kinase C (PKC) with phorbol esters upregulates α1A-a currents by ~10% ( Fig. 6c ). Because the domain I–II linker contributes to the PKC-dependent upregulation of some neuronal calcium channels34, we compared the modulation of α1A-a and α1A-b variants. Phorbol ester (PMA, 100 nM) caused twofold more upregulation of α1A-b channels than of α1A-a (+PMA α1A-a, 1.10 ± 0.01, +PMA α1A-b, 1.24 ± 0.03; Fig. 6c ). The presence of Val421 in the domain I–II linker was responsible for conferring the higher degree of upregulation associated with α1A-b channels ( Fig. 6d ). Overall, the results support the notion that the domain I–II linker is a crucial determinant for PKC and Gβγ modulation32 and suggest alternative splicing as a mechanism for controlling the second-messenger-dependent modulation of neuronal Ca2+ channels.

Discussion

Alternative splicing alters channel properties

Our present understanding of Ca2+ channel structure–function relationships primarily results from chimeric cDNA and mutagenesis studies, which have identified regions of Ca2+ channel α1 subunits important for permeation35, activation36 and inactivation37, excitation–contraction coupling38 and β subunit binding30. The functional effect of alternative splicing is an important tool because it provides insights into native amino acids essential for controlling channel properties.

Here we show that alternative splicing of the rat α1A gene results in variants with relatively small alterations in the domain I–II linker and domain S3–S4 regions. Two apparently distinct channel properties are affected by alternative splicing of N1605-P1606 in domain IV 53-54. The approximately 6-mV shift in current–voltage relationships associated with N1605-P1606 is consistent with studies on L-type Ca2+ channels showing that mutations both within and flanking S3 segments affect the voltage dependence of activation36. Many native Ca2+ channels display variable activation characteristics, which may generally result from alternative splicing in S3–S4 regions.

The second major change in channel properties associated with N1605-P1606 is to alter the sensitivity of α1A to the spider peptide toxin, ω-Aga IVA. ω-Aga IVA is a 48-residue positively charged peptide isolated from the venom of the funnel web spider Agelenopsis aperta, which has long been used as a diagnostic for native P/Q-type Ca2+ channels16,17. As in native P-type currents, block of both α1A-a and α1A-b variants by ω-Aga IVA in HEK cells was highly voltage sensitive and was reversed by application of positive prepulses. The major effects of the insertion of N1605-P1606 were to decrease kon and increase koff, which together decreased toxin affinity 11-fold. Both ω-Aga IVA and ω-CTx MVIIC act extracellularly to block P/Q-type channels, although by distinct blocking mechanisms; ω-CTx MVIIC acts as a traditional pore blocker (S.I. McDonough, I.M. Mintz & B.P. Bean Soc. Neurosci. Abstr. 21, 140.9, 1995) and ω-Aga IVA binds to the gating machinery to stabilize channel closed states39. In contrast to ω-Aga-IVA, the presence or absence of N1605-P1606 did not alter the block of α1A channels by ω-CTx MVIIC, which is consistent with these toxins' different mechanisms of action.

ω-Aga IVA belongs to a class of polypeptide neurotoxins that act via a common mechanism to alter the gating properties of voltage-gated ion channels. These include α and β scorpion toxins (Na+ channels)40,41, hanatoxin (K+ channels)42 and grammotoxin (Ca2+ channels)43. Our data are consistent with studies showing that peptide gating modifiers interact at a conserved structural gating modifier site that primarily includes S3–S4 loops. Single amino-acid changes in the domain II S3–S4 loop of Na+ channels and the S3–S4 loop of K+ channels significantly alter toxin affinities40,41,42,43. In this regard, the insertion of N1605-P1606 in α1A by splicing may alter the conformation of the domain S3–S4 loop and result in an unfavorable binding site for ω-Aga IVA.

The third change in α1A properties associated with alternative splicing is that the presence of Val421 dramatically slowed channel inactivation kinetics and affected steady-state inactivation. The Val421 residue is close to the site in the domain I–II linker that interacts with Ca2+ channel β subunits30. However, Val421 is unlikely to affect inactivation by disrupting β subunit binding because all β subunits still shifted α1A current–voltage properties. Furthermore, co-expression with the β2a subunit, which normally slows inactivation, caused an additional significant slowing of α1A-b currents. The presence of Val421 altered channel gating as reflected by these channels reopening throughout the depolarization. Assuming that Val421 alters the prevalence of different gating modes, these channels might enter an alternate mode with an increased rate constant for returning to the open state from the inactivated state. This less-stable inactivated state would be reflected in reopening during the depolarization.

The demonstration that the domain I–II linker contributes to Ca2+ channel inactivation44 ( Fig. 4 ) is distinct from studies of chimeric Ca2+ channels that suggested a predominant role for domain I in inactivation37. Although we did not detect any functional properties that could be attributed to alternative splicing of the EF hand region, it is unlikely that the specific splicing of this highly conserved structural motif is benign. Some possibilities include the interaction of this putative divalent ion binding site with intracellular cations, or the differential interaction at this site of yet-to-be identified proteins that bind to the Ca2+ channel carboxyl terminus.

Differential expression of α1A splice variants

P-type currents originally described in cerebellar Purkinje cells can be distinguished from Q-type currents in cerebellar granule cells by their higher sensitivity to ω-Aga-IVA and non-inactivating waveform15,16,17,18,19. We showed that α1A channels can exhibit the slower inactivation kinetics typically associated with P-type currents by associating with a co-expressed β2a subunit, possessing Val421 in the domain I–II linker, or both ( Fig. 2b ). In contrast, α1A channels lacking Val421 or associated with β1, β3 or β4 subunits exhibit the faster inactivation kinetics typically associated with Q-type currents ( Fig. 2b ). The β2a subunit is highly expressed in cerebellar Purkinje cells (data not shown); therefore the non-inactivating P-type currents may be encoded by the combination of α1A-a and β2a (and α2δ).

α1A transcripts both with and without N1605-P1606 in domain IV S3–S4 were expressed at approximately equal levels in the hippocampus. In contrast, most α1A transcripts in the cerebellum lacked N1605-P1606. The presence of variants that confer the lower Q-type and higher P-type affinities for ω-Aga IVA ( Fig. 5 ) predicts that both P-type and Q-type channels are widely expressed in the hippocampus, as reported in CA1-CA3 cells17,20. The data also predict that most α1A transcripts in the cerebellum would have the P-type pharmacological profile, as reported in Purkinje cells21. The more abundant cerebellar granule cells also highly express an inactivating Ca2+ current (called G1) with P-type pharmacology21, although some controversy remains regarding the nature of Q-type currents in granule cells. The Q-type current originally described in cultured cerebellar granule cells was reported to represent up to 35% of the whole-cell Ca2+ current19. However, others21 failed to identify a Q-type current in this same primary granule cell preparation, but instead described the G1 current. One possible explanation for these conflicting results is that variations in culture conditions differentially induce expression of α1A-a and α1A-b splice variants in cultured granule cells.

ω-Aga-IVA affinity and expression system considerations

In Xenopus oocytes, the currents from cloned α1A variants lacking N1605-P16061A-a) seem pharmacologically similar to native Q-type currents (Kd for ω-Aga-IVA > 100 nM; Fig. 5a )29,30. However, in HEK 293 cells, the Kd for ω-Aga IVA block of α1A-a was approximately 15 nM. In another study45, the identical rat brain α1A-a cDNA expressed in COS cells was found to have a Kd of ~11 nM. This 15–20-fold difference in ω-Aga IVA sensitivity between mammalian cells and Xenopus oocytes may reflect cell-type-specific proteins that interact directly or indirectly with α1A-a and alter pharmacological properties. Alternatively, cell-type-specific post-translational processing might alter channel affinity for ω-Aga IVA. For example, cloned acetycholine receptors (nAChR) expressed in Xenopus oocytes are incorrectly glycosylated compared to native nAChRs in Torpedo electroplax46. Because ω-Aga IVA is highly positively charged, the amounts and types of oligosaccharides (for example, high mannose versus complex) in the vicinity of the binding site may significantly affect toxin binding. Small differences in glycosylation patterns between mammalian cell lines and cerebellar Purkinje cells might also account for the difference in Kd between exogenously expressed α1A-a (Kd ≈ 11–15 nM) and native P-type currents (Kd ≈ 2 nM).

Alternative splicing: physiological consequences

It is important to gain an understanding of the contributions of α1A channels not only to normal neuronal functions, such as neurotransmitter release5,6,7,820, but also to pathological states such as migraine, ataxia and epilepsy27,28,29. Immunohistochemical staining localizes α1A calcium channels postsynaptically on cell bodies and dendrites as well as on presynaptic terminals47. In nerve terminals, both the 6-mV shift in α1A channel activation resulting from the splicing of N1605-P1606 and the slowed inactivation kinetics associated with Val421 would be predicted to alter the efficacy of synaptic transmission and significantly affect excitability.

Certain G-protein-coupled receptors inhibit P/Q-type and N-type channels by direct binding of Gβγ to the domain I–II linker of α1A and α1B subunits32. The alternatively spliced Val421 residue is within one of two Gβγ binding sites in the I–II linker, and its presence significantly increases slowing of the kinetics of G-protein-dependent inhibition and increases the overall degree of block ( Fig. 6a and b ). For both α1A-a and α1A-b, the kinetics of G-protein re-inhibition following a positive prepulse were well described by a single exponential. However, the time constant for re-inhibition was twice as long for α1A-b (and α1A-a + Val421) as for α1A-a. Within the few milliseconds of an action potential, the G-protein-inhibited α1A-b channels with their very slow activation ( Fig. 6a ) would be more effectively inhibited, and their ability to trigger fast neurotransmitter release might be attenuated compared with α1A-a. Similar to strong step depolarizations, bursts of action potentials can relieve G-protein inhibition48. Thus, the slower inhibition of α1A-b channels would be expected to result in a substantially larger amount of calcium influx compared to α1A-a (see Fig. 6b ) and might cause more pronounced synaptic efficiency for slow neurotransmitter release.

P-, Q- or other types?

The four different types of Ca2+ channel β subunits are widely expressed in the rat CNS and also have subtype-specific expression patterns. Because each of the β subunits can differentially modify α1A inactivation kinetics25, inactivation kinetics are a poor diagnostic of Ca2+ channel subtype without knowledge of biochemical composition. Similarly, we find significant differences in the sensitivity of α1A channels to ω-Aga IVA based on alternative splicing and the cell type used for expression. Thus, ω-Aga IVA affinity might not always be a reliable diagnostic of specific α1A variants.

It is important to emphasize that the electrophysiological and pharmacological properties of the α1A-a and α1A-b variants described here do not exactly match those of either prototypical P-type or Q-type currents. Indeed, α1A-a and α1A-b describe two types of Ca2+ channels that exhibit some characteristics of both P-type and Q-type currents. Native Ca2+ channels with properties intermediate between P- and Q-type have been described in neurons and endocrine cells21,22,49. For example, the cerebellar granule cell G1 Ca2+ channel is highly sensitive to ω-Aga-IVA (similar to P-type) yet exhibits substantial inactivation (similar to Q-type)21. We predict that G1 is likely to be encoded by a splice variant missing both Val421 and N1605-P1606 (for example, α1A-a).

The α1A-a and α1A-b variants examined here were each derived from single cDNAs; thus their properties likely reflect some portion of native α1A Ca2+ currents. In addition, other α1A variants are expressed that exhibit both some combination of the alternatively spliced regions examined here as well as other independent alternatively spliced sequences23,27,29,49,50. Assuming a mutually exclusive splicing mechanism, our data suggest that Q-type channels represent only one possible phenotypic variant resulting from splicing of a single α1A gene encoding the P-type Ca2+ channel.

Methods

Electrophysiological recording.

Preparation of Xenopus oocytes, nuclear injection with Ca2+ channel subunit cDNAs, two-microelectrode voltage clamp, and cell-attached single-channel, patch-clamp recordings were done as described25,34. For macroscopic currents, the bathing medium contained 40 mM Ba(OH)2, 25 mM TEA-OH, 25 mM NaOH, 2 mM CsOH and 5 mM HEPES (titrated to pH 7.3 with methane-sulfonic acid). During voltage-clamp recordings, oocytes were routinely injected with BAPTA to eliminate the endogenous oocyte Ca-dependent chloride current. For G-protein modulation, Ca2+ channels were co-expressed with the μ-opioid receptor and G-protein activation was triggered via application of 1 μM DAMGO. For PKC modulation, oocytes were perfused with 100 nM PMA as described34, and BAPTA was omitted.

For the determination of current–voltage relationships, curves were fitted according to the equation I = G × (V – E)/[1 + exp((V –V0.5)/k)], where I is the current at a given voltage V, G is the conductance, E is the extrapolated reversal potential, V0.5 is the voltage for half activation and k is the slope factor. For steady-state inactivation, cells were held at various potentials for 15 s and currents recorded during a subsequent test pulse to the peak potential of the I–V. Steady-state inactivation was calculated with a Boltzmann equation and the potential for half inactivation determined from the best fit. All values are given as mean ± s.e. Leak currents and capacitive transients were subtracted using a P/5 procedure.

Transient transfection of human embryonic kidney cells (HEK-tsa201) was done by standard calcium-phosphate precipitation, and whole-cell patch-clamp recordings were as described48. The external recording solution was 5 mM BaCl2, 1 mM MgCl2, 10 mM HEPES, 40 mM TEACl, 25 mM glucose and 65 mM CsCl (pH 7.2). The internal pipette solution was 105 mM CsCl, 25 mM TEACl, 1 mM CaCl2, 11 mM EGTA and 10 mM HEPES (pH 7.2). Peptide toxins were perfused onto the cells using a gravity-driven microperfusion system.

Molecular biology.

To identify the α1A-b cDNA, we screened a rat brain cDNA library24 with a [32P]deoxyoligonucleotide probe homologous to domain I of previously cloned neuronal α1 subunits. Of seventeen clones identified as α1A types, the complete DNA sequence of the single 7.2 kb α1A-b cDNA was determined on both strands. The α1A-b (–V) construct was made by substituting a 2.4-kb Xho I/Nhe I fragment of α1A-a into α1A-b. The α1A-a(+V) construct was made by substituting a 1.8-kb Tth III1 fragment of α1A-b into α1A-a. The α1A-a (+NP) cDNA was constructed by substituting a 2.5-kb BamHI fragment from α1A-b into α1A-a. The α1A-b (–NP) construct was created by switching the 2.5-kb fragment from α1A-a into α1A-b. Chimeric cDNAs were originally constructed in the plasmid Bluescript and then subcloned into the Xho I/Spe I sites of the nuclear expression vector, pMT2. Each construct was verified by DNA sequencing across the chimera junctions and through the altered region.

To examine domain I–II linker splice junctions, we screened a rat genomic library (lambda DASH, Stratagene) using a [32P]-labeled cDNA fragment to the α1A-a domain I–II linker. A 5.2-kb EcoRI fragment that hybridized to oligonucleotide probes specific for α1A sequences at positions 1200 and 1275 bp was selected and DNA sequencing used to identify the intron–exon boundaries across the I–II linker. Splicing in domain IV was analyzed by using PCR to amplify rat genomic sequences flanking S3 and S4. Direct sequencing of the PCR products identified the intron–exon boundaries.

For PCR detection of V421, a sense oligonucleotide common to both α1A-a and α1A-b variants was used in conjunction with an antisense oligonucleotide containing the antisense sequence for Val421 at its 3´ end. In situ localization was done using 30 μm paraformaldehyde-fixed sections from 250 to 300 gm adult male rat brains. Sections were hybridized overnight at 37°C in 50% formamide, 0.6 M NaCl, 1 mM EDTA, 100 mM DTT, 10 mM Tris pH 7.5, 1× Denhardts, 10% dextran sulphate, 0.1% SDS, 0.1% sodium thiosulphate, 0.05% denatured salmon sperm DNA and 0.05% yeast tRNA. Probes consisted of [α-35S]dATP end-labeled antisense 39-base synthetic deoxyoligonucleotides derived from the unique EF hand regions of α1A-a and α1A-b. Sections were washed at 50°C in 0.5× SSC, 1% sodium thiosulphate and 14 mM β-mercaptoethanol, exposed to X-ray film for 1 week, then dipped in Kodak photographic emulsion NTB-2 and exposed for 2–4 weeks.