The importance of voltage-gated Ca2+ channels in cellular function is illustrated by the many distinct types of Ca2+ currents found in vertebrate tissues, a variety that is generated in part by numerous genes encoding Ca2+ channel subunits. The degree to which this genetic diversity is shared by invertebrates has only recently become apparent. Cloning of Ca2+ channel subunits from various invertebrate species, combined with the wealth of information from the Caenorhabditis elegans genome, has clarified the organization and evolution of metazoan Ca2+ channel genes. Functional studies have employed novel structural information gained from invertebrate Ca2+ channels to complement ongoing research on mammalian Ca2+ currents, while demonstrating that the strict correspondence between pharmacological and molecular classes of vertebrate Ca2+ channels does not fully extend to invertebrate tissues. Molecular structures can now be combined with physiological data to develop a more cogent system of categorizing invertebrate channel subtypes. In this review, we examine recent progress in the characterization of invertebrate Ca2+ channel genes and its relevance to the diversity of invertebrate Ca2+ currents.

Voltage-gated Ca2+ channels perform a unique role in excitable cells. By gating Ca2+ influx in response to depolarization, they couple changes in membrane potential to numerous intracellular mechanisms regulated by Ca2+, including neurotransmitter release, muscle contraction, gene expression and cellular differentiation. Many of these responses are tuned to a specific temporal and spatial pattern of Ca2+ entry (e.g. Gu and Spitzer, 1995; Dolmetsch et al., 1997), which must be balanced with the potential toxicity caused by high intracellular levels of Ca2+. The need for carefully regulated patterns of Ca2+ influx in vertebrate cells is reflected by the presence of several molecular subtypes of Ca2+ channels, each distinct in its kinetics, pharmacology and tissue distribution. Physiological evidence suggests that invertebrate tissues also possess a variety of Ca2+ currents, but the molecular diversity of the channels gating these currents has not been as thoroughly studied.

A number of recent advances have now directed greater attention towards invertebrate Ca2+ channels. Foremost among these was the virtual completion of the Caenorhabditis elegans genome sequence in late 1998 (The C. elegans Sequencing Consortium, 1998). For the first time, we can examine the complete array of genes required for an organism possessing a nervous system, an approach that is not yet possible for a vertebrate model. Furthermore, novel C. elegans genes have led directly to the identification of new families of channel subunits (e.g. Wei et al., 1996), including the long-anticipated discovery of mammalian genes encoding low-voltage-activated T-type Ca2+ channels (Perez-Reyes et al., 1998). Progress has extended to other invertebrate models as well. Ca2+ channel subunits have been cloned from organisms in several phyla, providing a broader view of the evolution of channel genes. Functional expression of cloned subunits (Jeziorski et al., 1998; Okamura, 1999) now allows direct comparison between the molecular and physiological properties of invertebrate channels. Finally, classical genetic approaches in Drosophila melanogaster (Zheng et al., 1995; Smith et al., 1996, 1998) and C. elegans (Schafer and Kenyon, 1995; Lee et al., 1997) have correlated numerous channel mutations with their physiological effects.

The value of these findings extends well beyond the study of invertebrate physiology. In addition to elucidating how channels evolved, genes from early metazoans provide a perspective on channel structure and function that is unavailable from mammalian studies. This review summarizes the latest work on the molecular biology of invertebrate Ca2+ channels, with particular emphasis on their structural heterogeneity and their relationship to mammalian channel subtypes. The reader is also referred to thorough reviews of the physiology of invertebrate Ca2+ currents (Kits and Mansvelder, 1996; Skeer et al., 1996), the molecular biology of vertebrate Ca2+ channels (Stea et al., 1995; Catterall, 1995; Perez-Reyes and Schneider, 1995; De Waard et al., 1996; Jones, 1998) and the role of auxiliary subunits in channel function (Walker and De Waard, 1998).

Although Ca2+ currents were first identified in invertebrate tissues (Fatt and Ginsborg, 1958), the pharmacological and molecular identification of channel subtypes has been based primarily on studies in vertebrate preparations. Because the resulting criteria are commonly used to define invertebrate channels as well, a brief overview of vertebrate channel physiology is necessary to understand the present classification of invertebrate currents.

The several types of vertebrate Ca2+ currents that have been identified fall into two categories, high-voltage-activated (HVA) and low-voltage-activated (LVA). HVA currents can be further subdivided into L-type currents, which are sensitive to blockage or modulation by dihydropyridines (DHPs), and the various DHP-insensitive currents collectively known as non-L-type. Each current is gated by a channel that includes a large, pore-forming α1 subunit; HVA channels also contain several auxiliary subunits that have not, as yet, been identified in LVA channels. Because the α1 subunit largely determines the physiology and pharmacology of the resulting current, each of the vertebrate α1 subunits that have been cloned and functionally expressed can be correlated with a previously defined Ca2+ current subtype. Three L-type α1 subunits (α1S, α1C, α1D) gate DHP-sensitive currents (Tanabe et al., 1987; Mikami et al., 1989; Hui et al., 1991; Williams et al., 1992b). A fourth L-type subunit recently cloned from human retina, α1F (Bech-Hansen et al., 1998; Strom et al., 1998), has not yet been expressed, but is grouped with the L-type channels on the basis of molecular similarity. Non-L-type α1 subunits can be further distinguished by their responses to various peptide toxins (for a review, see Olivera et al., 1994). The α1B subunit (Williams et al., 1992a; Dubel et al., 1992) gates the N-type current that is blocked by the snail toxin ω-conotoxin GVIA (ω-CTx GVIA), while α1A (Mori et al., 1991; Starr et al., 1991) is responsible for the P/Q-type current that is blocked by ω-conotoxin MVIIC (ω-CTx MVIIC) and the spider toxin ω-agatoxin IVA (ω-Aga IVA). A third non-L-type subunit, α1E (Niidome et al., 1992; Soong et al., 1993), gates the toxin-resistant R-type current (Zhang et al., 1993; Randall and Tsien, 1997). LVA, or T-type, α1 subunits have only recently been cloned, but three subtypes are already known: α1G (Perez-Reyes et al., 1998), α1H (Cribbs et al., 1998) and α1I (Lee et al., 1999b). L-type, non-L-type and T-type α1 subunit genes form three phylogenetic clusters, such that proteins within a given class are significantly more similar than are those from separate classes (see Figs 1, 2).

Fig. 1.

Dendrogram of α1 subunits. For the purposes of this comparison, only the relatively conserved region spanning domains III and IV of each sequence was used. The boundaries of the domains for the more distant molecules were estimated after an initial alignment. A subsequent alignment was completed using ClustalW (Thompson et al., 1994), and the neighbor-joining algorithm (Saitou and Nei, 1987) was then used to construct the tree. Comparable results were obtained using related algorithms. Invertebrate subunits are colored in red. The clusters corresponding to the three known classes of voltage-gated Ca2+ channels are colored separately. Accession numbers are as follows (see Table 1 and text for others): jellyfish Na+ (L15445); human α1S (A55645); rabbit α1S (M23919); carp α1S (A37860); Fugu α1S (AAC15583); human α1C (Z74996); rabbit α1C (X15539); human α1D (M76558); rat α1D (M57682); human α1F (AJ224874); human α1A (X99897); rat α1A (M64373); human α1B (M94172); rat α1B (M92095); ray α1B (L12532); human α1E (A54972); rat α1E (L15453); ray α1E (L12531); rat α1G (AF027984); human α1H (AF051946); rat α1I (AF086827); rat Rb21 (AF078779); yeast Ca2+ (P50077); yeast Na+ (Z98981).

Fig. 1.

Dendrogram of α1 subunits. For the purposes of this comparison, only the relatively conserved region spanning domains III and IV of each sequence was used. The boundaries of the domains for the more distant molecules were estimated after an initial alignment. A subsequent alignment was completed using ClustalW (Thompson et al., 1994), and the neighbor-joining algorithm (Saitou and Nei, 1987) was then used to construct the tree. Comparable results were obtained using related algorithms. Invertebrate subunits are colored in red. The clusters corresponding to the three known classes of voltage-gated Ca2+ channels are colored separately. Accession numbers are as follows (see Table 1 and text for others): jellyfish Na+ (L15445); human α1S (A55645); rabbit α1S (M23919); carp α1S (A37860); Fugu α1S (AAC15583); human α1C (Z74996); rabbit α1C (X15539); human α1D (M76558); rat α1D (M57682); human α1F (AJ224874); human α1A (X99897); rat α1A (M64373); human α1B (M94172); rat α1B (M92095); ray α1B (L12532); human α1E (A54972); rat α1E (L15453); ray α1E (L12531); rat α1G (AF027984); human α1H (AF051946); rat α1I (AF086827); rat Rb21 (AF078779); yeast Ca2+ (P50077); yeast Na+ (Z98981).

Fig. 2.

Sequence similarity among identified Ca2+ channel α1 subunits. As in Fig. 1, the region encompassing domains III and IV (approximately 25–35 % of the full length) was selected for each molecule, and the pairwise percentage of identical residues was determined after a global alignment. The upper right half of the figure gives the values for individual comparisons, while the lower left half presents mean values between the indicated groups of α1 subunits. In the latter instance, the means for comparisons within vertebrate types (e.g. L-type versus L-type) exclude within-subtype values (e.g. α1S versus α1S). The lengths of the region used are indicated for each molecule. Percentage identity was determined by dividing the number of identical residues by the length of a sequence; because the lengths of two molecules in a given comparison may differ slightly, the percentage may vary by 1–2 % depending on which length is used. Where several mammalian sequences for a given subtype were available, only two were included because of their relatively high level of conservation. Invertebrate sequences are shaded in gray. AAs, number of amino acid residues; Inver., invertebrate; Ver., vertebrate.

Fig. 2.

Sequence similarity among identified Ca2+ channel α1 subunits. As in Fig. 1, the region encompassing domains III and IV (approximately 25–35 % of the full length) was selected for each molecule, and the pairwise percentage of identical residues was determined after a global alignment. The upper right half of the figure gives the values for individual comparisons, while the lower left half presents mean values between the indicated groups of α1 subunits. In the latter instance, the means for comparisons within vertebrate types (e.g. L-type versus L-type) exclude within-subtype values (e.g. α1S versus α1S). The lengths of the region used are indicated for each molecule. Percentage identity was determined by dividing the number of identical residues by the length of a sequence; because the lengths of two molecules in a given comparison may differ slightly, the percentage may vary by 1–2 % depending on which length is used. Where several mammalian sequences for a given subtype were available, only two were included because of their relatively high level of conservation. Invertebrate sequences are shaded in gray. AAs, number of amino acid residues; Inver., invertebrate; Ver., vertebrate.

The various α1 subunits differ in both their tissue and subcellular distribution. The α1S subunit is found in skeletal muscle, and the α1C subunit in heart and other tissues, while the α1D and the non-L-type α1 subunits are confined primarily to brain. Within neurons, α1A and α1B are localized to synaptic terminals (Westenbroek et al., 1992, 1995), where they gate the Ca2+ influx necessary for transmitter release, whereas α1E, along with neuronal L-type channels, is found in the soma and proximal dendrites (Hell et al., 1993; Yokoyama et al., 1995).

Table 1.

Known invertebrate Ca2+ channel subunit sequences

Known invertebrate Ca2+ channel subunit sequences
Known invertebrate Ca2+ channel subunit sequences

Ca2+ channels are part of a superfamily of voltage-gated cation channels that also includes Na+ and K+ channels. The pore of a cation channel is surrounded by four homologous domains, which are contained within the single α1 subunit in Ca2+ channels and α subunit in Na+ channels, but are formed by four separate subunits in K+ channels. Each domain comprises six transmembrane segments, denoted S1–S6; the S4 segments, which contain positively charged residues at every third or fourth position, are believed to function as voltage sensors. The external loop between S5 and S6 of each domain projects into the lumen of the channel to form part of the pore. Within each pore region in HVA α1 subunits, a glutamate residue (E in single letter code) is found at a conserved position. Together, the four glutamates (EEEE) form a selectivity gate (Heinemann et al., 1992; Yang et al., 1993); the pore of LVA α1 subunits differs only slightly, with aspartates in place of glutamates at two sites (EEDD). Non-conservative substitutions for glutamate, such as non-polar alanine (A) or basic lysine (K) residues, decrease the selectivity of the channel for divalent over monovalent cations (Ellinor et al., 1995), as reflected by the DEKA motif found at the corresponding sites in most cloned Na+ channels. Thus, the presence of acidic residues in the four key positions is regarded as essential to a Ca2+-selective channel. The intracellular domains of HVA α1 subunits also contain highly conserved regions that are implicated in protein interaction. The Ca2+ channel β subunit binds to a specific region between domains I and II (Pragnell et al., 1994) that is well conserved among all vertebrate HVA α1 subunits. A portion of the II–III domain of the skeletal muscle α1S subunit is implicated in excitation–contraction coupling (Tanabe et al., 1990; Nakai et al., 1998; Grabner et al., 1999), while a segment of the I–II loop of non-L-type α1 subunits binds Gβγ subunits (De Waard et al., 1997; Zamponi et al., 1997), and calmodulin mediates Ca2+-dependent inactivation in L-type channels via a site in the proximal carboxyl tail (Zühlke et al., 1999). The α1 subunit also contains the binding sites for DHPs (Tang et al., 1993; Grabner et al., 1996) and other L-type channel blockers (Hockerman et al., 1995; Kraus et al., 1998) as well as peptide toxins (Ellinor et al., 1994). Conservation among mammalian α1 subunits is greatest within the transmembrane segments, the pore-forming loops and the intracellular interaction domains.

While the α1 subunit essentially determines the type of current gated, the auxiliary subunits have important modulatory roles in HVA channels. The β subunit aids in trafficking of the α1 subunit to the plasma membrane (Chien et al., 1995; Tareilus et al., 1997) and also influences the kinetics of the current, leading to activation at more hyperpolarized potentials as well as modulation of the rate- and voltage-dependence of inactivation (Wei et al., 1991; Lacerda et al., 1991; Varadi et al., 1991). The four β subtypes (β1–β4) that have been cloned from mammals thus far (Ruth et al., 1989; Perez-Reyes et al., 1992; Castellano et al., 1993a,b) differ in their effects upon α1 subunits (Olcese et al., 1994). Strong conservation within the interaction domains of α1 and β subunits allows several β subtypes to associate with native α1A (Liu et al., 1996), α1B (Scott et al., 1996) and neuronal L-type α1 subunits (Pichler et al., 1997), leading to potential heterogeneity in the resulting currents. The α2δ subunit (Ellis et al., 1988) consists of the heavily glycosylated extracellular α2 domain linked by disulfide bonds to the membrane-spanning δ segment, both derived from the same gene. The α2δ subunit also modulates the α1 subunit (Singer et al., 1991), although less dramatically than does the β subunit. A second and third mammalian subtype of the α2δ subunit have recently been reported (Klugbauer et al., 1999). The γsubunit (Jay et al., 1990) was once thought to be present only in skeletal muscle Ca2+ channels. Recently, a γ homologue, named stargazin on the basis of its associated mutation in mice, has been discovered in brain (Letts et al., 1998) and is able to modulate the inactivation of heterologously expressed neuronal α1 subunits.

A comprehensive survey of the physiology of invertebrate Ca2+ currents is beyond the scope of this review. Instead, we will attempt to summarize certain findings that consistently emerge from the literature (for reviews, see also Kits and Mansvelder, 1996; Skeer et al., 1996). There is considerable electrophysiological and pharmacological evidence for diversity of invertebrate Ca2+ currents (e.g. Pelzer et al., 1989). Both LVA and HVA currents have been recorded from neurons (Haydon and Man-Son-Hing, 1988; Yamoah and Crow, 1994) and other invertebrate tissues (Hagiwara et al., 1975; Arnoult and Villaz, 1994). Evidence for more than one neuronal LVA current in cockroach (Grolleau and Lapied, 1996) and mollusc (Kits and Mansvelder, 1996) has been reported. Similarly, invertebrate HVA currents may be separated into components distinguished by kinetics or by responses to channel blockers. Mammalian L-type currents are blocked by three major classes of L-type channel ligands, DHPs, phenylalkylamines (PAAs) and benzothiazepines, that bind to distinct but overlapping sites on mammalian L-type channels (for a review, see Hockerman et al., 1997). The potency of these ligand classes in blocking invertebrate Ca2+ currents appears to correlate less closely than in vertebrate tissues. PAAs are more potent than DHPs in blocking neuronal current components in Drosophila (Pelzer et al., 1989; Morales et al., 1999), cockroach (Mills and Pitman, 1997; Wicher and Penzlin, 1997) and honeybee (Schäfer et al., 1994), whereas DHPs such as nifedipine are more effective in Drosophila muscle (Gielow et al., 1995) and Lymnaea neurons (Dreijer and Kits, 1995) and also block currents in crayfish muscle (Araque et al., 1994) and Aplysia tissues (Nerbonne and Gurney, 1987; Brezina et al., 1994; Laurienti and Blankenship, 1997; Fieber, 1998). The DHP agonist BayK 8644 stimulates a current in crustacean muscle (Erxleben and Rathmeyer, 1997), but not the nifedipine-sensitive current in Aplysia neurons (Nerbonne and Gurney, 1987; however, see Trudeau et al., 1993). Of the HVA currents that are DHP-insensitive, certain components can be blocked by peptide toxins. A current in crayfish neurons is blocked by ω-Aga IVA, but not by ω-CTx GVIA or L-type ligands, leading to its identification as P-type (Araque et al., 1994; Wright et al., 1996; Hong and Lnenicka, 1997; García-Colunga et al., 1999); ω-Aga IVA also specifically blocks Ca2+ current components in squid neurons (Llinas et al., 1989; McFarlane and Gilly, 1996), locust neurons (Bickmeyer et al., 1994) and sea urchin eggs (Vogel et al., 1999). In neurons from cockroach (Wicher and Penzlin, 1997) and Aplysia (Fossier et al., 1994), ω-Aga IVA and ω-CTx GVIA appear to block distinct HVA current components, suggesting multiple non-L-type channels. In certain instances, a neuronal HVA current is insensitive to both L-type and non-L-type ligands (Richmond et al., 1996; Hong and Lnenicka, 1997).

These studies, as well as some reviews (Olivera et al., 1994; Skeer et al., 1996), have frequently concluded that invertebrate Ca2+ currents do not correspond precisely to the profiles of defined vertebrate current subtypes. While invertebrate currents are often identified as L-type, non-L-type or T-type, many authors emphasize the significant biophysical or pharmacological differences from vertebrate currents. Furthermore, an invertebrate Ca2+ current that does resemble a particular vertebrate class may not necessarily be gated by a channel of the corresponding molecular subtype. A possible example is the current in arthropod neurons described as P-type on the basis of its sensitivity to ω-Aga IVA, but not to ω-CTx GVIA. The greater potency in invertebrate tissues of ω-Aga IVA and other spider toxins (Llinas et al., 1989; Leung et al., 1989) is not surprising, because ω-CTx GVIA and other toxins generated by fish-hunting snails of the genus Conus are designed for optimal effect in vertebrate tissues, whereas spider toxins are targeted towards insects. Thus, the invertebrate currents blocked by ω-Aga IVA, a toxin that binds to the α1A subunit in mammalian tissues, may be gated by α1 subunits of less specific molecular identity that are nonetheless sensitive to certain toxins.

For these reasons, invertebrate Ca2+ channels are best examined by a combination of physiological and molecular approaches. Studies in vertebrate tissues have benefited from such a synthesis, demonstrating that a wide range of currents can result from small variations in channel structure. Given the evidence for comparable current heterogeneity in invertebrate tissues, the composition of invertebrate Ca2+ channels is of paramount interest.

In contrast to other classes of channels, such as voltage-gated K+ channels (Papazian et al., 1987), that were first cloned from invertebrates, no invertebrate Ca2+ channel genes were identified until several mammalian channel subunits had already been sequenced and expressed, and a classification framework had been established. Recent progress in the cloning of invertebrate Ca2+ channels has been rapid, however (Table 1). To date, full-length L-type homologues have been isolated from the ascidian Halocynthia roretzi (Okamura, 1999), the arthropods Drosophila melanogaster (Zheng et al., 1995) and Musca domestica (Grabner et al., 1994a), the nematode C. elegans (Lee et al., 1997) and the cnidarians Cyanea capillata (Jeziorski et al., 1998) and Stylophora pistillata (Zoccola et al., 1999). Non-L-type α1 subunits have been cloned from Drosophila (Smith et al., 1996) and the mollusc Loligo bleekeri (Kimura et al., 1997), as well as a nearly complete sequence from C. elegans (Schafer and Kenyon, 1995). In addition, the coexistence of L-type and non-L-type channels has been established in the arthropod Blattella germanica (Inagaki et al., 1998), the mollusc Aplysia californica (White and Kaczmarek, 1997) and the platyhelminth Bdelloura candida (Greenberg et al., 1995) by the amplification of multiple cDNA fragments. (No invertebrate T-type α1 subunits, with the exception of the C. elegans gene discussed in the following section, have yet been reported.) As is true for other classes of voltage-gated and ligand-gated ion channels, Ca2+ channel α1 subunits are well conserved between vertebrates and invertebrates. Similarity is highest in the transmembrane domains and in the pore regions, where the EEEE motif is invariably found. The extended intracellular domains tend to vary more, except in regions that have been implicated in channel regulation, such as the I–II loop and the proximal carboxyl tail.

Examination of invertebrate Ca2+ channel genes, either by phylogenetic analysis (Fig. 1) or by direct measurement of amino acid identity (Fig. 2), reveals numerous clues to the pattern of channel evolution. Each α1 subunit that has been cloned from an invertebrate can be unambiguously identified as either L-type or non-L-type, but is not aligned with a specific mammalian subtype within the classes (Fig. 1), suggesting that the various subtypes arose during the evolution of vertebrates. For instance, the Drosophila subunits initially identified as α1A and α1D are now believed to have appeared well before the divergence of individual subtypes (Smith et al., 1996; Peixoto et al., 1997). Subunits from different classes are no more similar to each other in early invertebrates than in mammals, indicating that the separation of channel families occurred very early in, or prior to, the emergence of metazoans. In addition, no convergence is evident between invertebrate Ca2+ channels and Na+ channels, even in the jellyfish Cyanea, a member of the earliest extant phylum in which a nervous system is expressed.

No Ca2+ channel gene has yet been unambiguously identified in any organism predating metazoans. However, Cch1, a cloned yeast protein that has been identified as a homologue of Ca2+ channel α1 subunits (Paidhungat and Garrett, 1997), may have evolved from an ancestral Ca2+ channel. The predicted sequence of Cch1 displays certain hallmarks of the pore-forming subunits of both Ca2+ and Na+ channels, including a four-domain structure, similar membrane topology and positively charged residues within the S4 domains. Significantly, the glutamate residues conserved in the pore of HVA Ca2+ channels are found at appropriate positions within three of the four pore loops of Cch1, with an asparagine present in domain I (NEEE). Replacement of the corresponding glutamate in domain I of the rabbit α1C subunit with glutamine, which is chemically similar to asparagine, diminishes the affinity of the pore for Ca2+ only twofold (Ellinor et al., 1995). Yeast strains carrying a mutant form of the Cch1 gene exhibit decreased levels of intracellular Ca2+ (Paidhungat and Garrett, 1997), but the protein has not yet been directly shown to respond to voltage or to gate Ca2+ influx selectively. A second yeast gene also encodes a four-domain protein of unknown function that has been labeled a Na+ channel homologue. However, the two yeast proteins share the NEEE motif in the pore region and are more similar to each other than to any other member of the respective ion channel families; neither subunit exhibits preferential similarity to either Ca2+ or Na+ channels (Fig. 2).

Information regarding auxiliary subunits of invertebrate Ca2+ channels is not as extensive as that for α1 subunits. β subunits have been cloned from Musca (Grabner et al., 1994b) and Cyanea (Jeziorski et al., 1999). Neither protein is specifically similar to any of the four mammalian β isoforms; thus, both diverged from an ancestral β subunit prior to the appearance of the mammalian subtypes (Fig. 3). The Musca β subunit, when expressed in oocytes, increases the amplitude of the endogenous non-L-type Ca2+ current and also increases DHP binding to COS7 cells expressing the mammalian α1C subunit (Grabner et al., 1994b). The Cyanea β subunit modulates the cloned Cyanea L-type α1 subunit in a manner similar to that reported for mammalian α1–β interactions. Coexpression of the β subunit in oocytes increases the current produced by the α1 subunit, shifts its voltage-dependence to more negative potentials and accelerates its rate of inactivation (Jeziorski et al., 1999).

Fig. 3.

Dendrograms of α2δ and β subunits. (A) Dendrogram of Caenorhabditis elegans and selected mammalian α2δ subunits. Because the identified coding region of the unc-36 gene from C. elegans is apparently terminated before the true carboxyl terminus, the other sequences were truncated at an analogous position to facilitate the comparison. Accession numbers (see Table 1 and text for others): rabbit (P13806); pig (AF077665); human (AF042792); mouse (AJ010949). (B) Dendrogram of published invertebrate and selected vertebrate β subunit sequences, as well as the second putative β subunit homologue (w10c8.1) from C. elegans. Accession numbers (see Table 1 and text for others): human β1 (NP_000714); human β3 (NP_000716); Xenopus laevis β3 (AAA75519); Fugu rubripes β3 (U72484); human β4 (NM_000726); human β2 (NP_000715).

Fig. 3.

Dendrograms of α2δ and β subunits. (A) Dendrogram of Caenorhabditis elegans and selected mammalian α2δ subunits. Because the identified coding region of the unc-36 gene from C. elegans is apparently terminated before the true carboxyl terminus, the other sequences were truncated at an analogous position to facilitate the comparison. Accession numbers (see Table 1 and text for others): rabbit (P13806); pig (AF077665); human (AF042792); mouse (AJ010949). (B) Dendrogram of published invertebrate and selected vertebrate β subunit sequences, as well as the second putative β subunit homologue (w10c8.1) from C. elegans. Accession numbers (see Table 1 and text for others): human β1 (NP_000714); human β3 (NP_000716); Xenopus laevis β3 (AAA75519); Fugu rubripes β3 (U72484); human β4 (NM_000726); human β2 (NP_000715).

An advantage of invertebrate models such as Drosophila and C. elegans is their genetic tractability, which has revealed that mutations in Ca2+ channel genes are responsible for a variety of physiological defects. Mutant alleles have been identified for both known α1 subunit genes in Drosophila. The non-L-type gene maps to loci implicated in visual impairment (nightblind-A) and altered courtship song (cacophony) as well as a lethal variant (Smith et al., 1996). Two mutations of the L-type subunit have been characterized at the genetic level (Eberl et al., 1998). One, a lethal form, is truncated after the IVS4 segment. A second, partial loss-of-function phenotype results from a single nonconservative substitution within the IS1 segment, which slows activation of the channel (Ren et al., 1998). Changing the corresponding residue in a mammalian channel similarly alters its kinetics, establishing the importance of the IS1 segment in channel activation. Mutations have also been identified in two C. elegans α1 subunits. The unc-2 mutation in C. elegans lies in a gene encoding a non-L-type α1 subunit (Schafer and Kenyon, 1995). The product of the unc-2 allele lacks the IVS2 transmembrane segment, probably leading to a severely altered topology and a nonfunctional channel. Animals homozygous for unc-2 do not desensitize to the inhibitory effects of dopamine upon motor activity and the stimulation of egg-laying induced by serotonin, which implies a role for the intact channel in modulating neurotransmitter release. The C. elegans egl-19 mutation is found in a gene encoding an L-type channel α1 subunit (Lee et al., 1997). At least 26 mutations, representing myotonic (gain-of-function), flaccid (reduction-of-function) and lethal (loss-of-function) mutants, are localized to the egl-19 gene. Of the gain-of-function alleles that have been identified, two contain point mutations in or near the IS6 segment and two encode conservative changes in IIIS4. Each alteration is thought to affect the rate of channel inactivation, leading to extended depolarization of muscle cells.

The tissue distribution of invertebrate channel subtypes is generally similar to that in mammals, with L-type subunits expressed in both muscle and neural tissue, but non-L-type α1 subunits localized primarily to neurons. An example is seen in Aplysia, in which both types are present in neurons, but only the L-type is found in muscle. Within ganglia, the L-type channel is uniformly distributed throughout neurons and glia, while the non-L-type channel is restricted to somata, processes and growth cones, exhibiting a punctate labeling that suggests localization in vesicle membranes (White and Kaczmarek, 1997). In C. elegans, levels of expression of the L-type (Lee et al., 1997) and non-L-type (Schafer and Kenyon, 1995) α1 subunits are higher in muscle and neurons, respectively, although each transcript is found in both cell types. Transcripts encoding both the L-type (Zheng et al., 1995) and non-L-type α1 subunits (Smith et al., 1996) in Drosophila embryo are found primarily in brain, but functional studies have demonstrated that the L-type channel gates the majority of the Ca2+ current in larval muscle (Ren et al., 1998); the L-type transcript is more widely expressed in the adult (Zheng et al., 1995). The Musca L-type channel transcript is detected in larval muscle, but not neurons (Grabner et al., 1994a), while the ascidian L-type α1 subunit is found in both neural and muscle tissue (Okamura, 1999). In contrast, the non-L-type channel from squid is expressed throughout the nervous system, but not in heart, muscle or stomach (Kimura et al., 1997).

Functional expression of cloned invertebrate α1 subunits has allowed a direct assessment of the relationship between the molecular identity and pharmacology of an invertebrate channel. The sequence of the α1 subunit isolated from Cyanea (Jeziorski et al., 1998) is significantly more similar to that of mammalian L-type α1 subunits than to that of non-L-type or T-type subunits. Nevertheless, expression of the Cyanea channel in Xenopus laevis oocytes produces a rapidly inactivating current that is insensitive to L-type channel ligands as well as to peptide toxins. DHPs and PAAs that fully block vertebrate L-type currents at low micromolar concentrations have only weak effects upon the Cyanea channel (<50 % inhibition by each agent at 100 μmol l−1), and BayK 8644 fails to potentiate the current. The L-type α1 subunit from the tunicate Halocynthia is similarly insensitive to block or enhancement by DHPs when expressed in oocytes (Okamura, 1999). These findings are in accord with the indirect measurement of the DHP sensitivity of the Musca L-type α1 subunit. Three domains of mammalian L-type α1 subunits (IIIS5, IIIS6, IVS6) have been implicated in forming the DHP binding site; transfer of these domains from the Musca channel to the DHP-insensitive mammalian α1A subunit resulted in a current that could be blocked by the DHP isradipine (65 % inhibition at 10 μmol l−1), but was unstimulated by BayK 8644 (Sinnegger et al., 1997).

The above results demonstrate that the close association between the pharmacological and molecular identities of mammalian α1 subunits is not maintained in invertebrate channels. Such a distinction can be problematic. L-type currents, for example, are classically defined by their sensitivity to DHPs, so the description of a DHP-insensitive L-type channel may be viewed as contradictory. However, the similarity of the Cyanea and Halocynthia α1 subunits to mammalian L-type channels is approximately 10 % greater than that to non-L-type channels, whereas the DHP-binding pocket in L-type α1 subunits is formed by only nine residues, less than 1 % of the molecule (Mitterdorfer et al., 1996; Peterson et al., 1996; Sinnegger et al., 1997). Furthermore, the sites to which DHPs or toxins bind have no known endogenous ligands, implying that the formation of such sites is incidental to channel function. We therefore propose that the molecular definition of channel subtype is the more rigorous one, and that invertebrate currents that have previously been defined by pharmacological criteria may be reclassified once the peptide sequence of the corresponding channel is known.

The sequencing of the C. elegans genome provides an unprecedented view into the molecular neurobiology of an organism. The genes within an animal possessing a nervous system and other differentiated tissues can now be examined in full; for this reason, the Ca2+ channel subunit homologues identified in C. elegans are considered here separately. Furthermore, the C. elegans genome project has led to the discovery of novel genes that could not be detected using conventional homology-based cloning techniques (for a review, see Bargmann, 1998). Two algorithms, TBLASTN (Altschul et al., 1990), which searches the genome using reverse translation of a peptide sequence, and Genefinder (P. Green and L. Hillier, unpublished), which detects potential exons on the basis of consensus boundaries in C. elegans, have greatly aided the identification of putative channel genes.

Five C. elegans genes encode proteins that exhibit structural similarity to Ca2+ channel α1 subunits. Two of these, the egl-19 and unc-2 genes noted above, have been independently cloned by groups investigating mutant C. elegans strains (Schafer and Kenyon, 1995; Lee et al., 1997) and produce channel subunits corresponding to the L-type and non-L-type classes, respectively. The third α1 subunit gene in the C. elegans genome was first found in the c54d2 cosmid and is labeled c54d2.5. Although the function of c54d2.5 was unknown at the time of its discovery, the subsequent cloning of three mammalian homologues, each of which encodes a T-type channel (Perez-Reyes et al., 1998; Cribbs et al., 1998; Lee et al., 1999b), implies that the C. elegans subunit also forms an LVA channel. The reported open reading frame of c54d2.5, based on detection of potential exons, probably lacks at least one exon comprising domains IIIS3–IIIS4 (a candidate exon for this region is found between bases 29 066 and 28 825 on the antisense strand of c54d2). Elements that appear to be specific to the primary structure of LVA channels, such as an EEDD ring within the pore, are also found in the c54d2.5 peptide sequence. Significantly, egl-19, unc-2 and c54d2.5 are the sole C. elegans genes that fall within the respective L-type, non-L-type and T-type classes.

Two additional C. elegans genes encode novel proteins that exhibit distant similarity to both Ca2+ and Na+ channel subunits. One gene, comprising an open reading frame with a proposed length of 1410 amino acid residues, is found in the c27f2 cosmid (c27f2.3; accession number AAA81424), and the other, encoding a protein of 1581 residues, lies within c11d2 (c11d2.6; accession number AAC02575). The two proteins are equally similar to the pore-forming subunits of Ca2+ and Na+ channels, but display somewhat greater resemblance to one another. The sequences of the two nematode genes were used to clone a rat homologue of unknown function, named Rb21, that is expressed in brain and heart (Lee et al., 1999a); a short homologous fragment has also been found in the Schistosoma mansoni genome (accession number AQ400292). Both the C. elegans and Rb21 proteins have the four-domain structure common to voltage-gated Ca2+ and Na+ channels. The regions corresponding to the S4 segments possess some, but not all, of the positively charged amino acids characteristic of voltage-gated channels, and the key residues within the pore regions at positions that influence ionic selectivity (EEKE) differ from those in both Ca2+ (EEEE in HVA subunits or EEDD in LVA subunits) and Na+ (DEKA in vertebrate subunits) channels. We have cloned the cDNA corresponding to the c27f2.3 gene product from a C. elegans cDNA library (M. C. Jeziorski, unpublished results). The cDNA includes four additional exons beyond the 3′ end of the reported sequence, extending the open reading frame from 1410 to 1687 codons. However, the cloned 5′ end, which matches the sequence proposed for the cDNA in the C. elegans database, appears to lack a segment encoding the N terminus, such that the open reading frame is initiated within the S1 segment of domain I. Efforts to obtain an alternative clone for the 5′ end have not yet succeeded and, like Rb21 (Lee et al., 1999a), the c27f2.3 product fails to generate a voltage-gated cation current when expressed in oocytes (M. C. Jeziorski, unpublished results).

Two genes that encode apparent homologues of Ca2+ channel β subunits are also found within the C. elegans genome. One forms a protein that is strongly similar to that of known β subunits (55 % identical to the human β3 subunit) and is highly conserved in the primary interaction site. Because the protein closely resembles other invertebrate β subunits that have been cloned and expressed, it is likely to share with them the modulatory effects upon HVA α1 subunits that are typical of β subunits. The second C. elegans homologue, w10c8.1 (accession number AAB97592), is far less similar to β subunits (22 % identical to β3), but exhibits limited identity within the interaction domain and other regions that are conserved among cloned β subunits. The w10c8.1 cDNA, which we have isolated from a C. elegans library, contains an additional unreported exon of 132 bases (corresponding to bases 10 872–10 741 on the antisense strand of C. elegans cosmid w10c8) that extends the open reading frame to 357 residues, approximately 80 % of the average length of conventional β subunits. The function of w10c8.1 remains unresolved, however. Coinjection of Xenopus laevis oocytes with RNA encoding w10c8.1 fails to modulate the expressed Cyanea L-type α1 subunit, a channel that can be regulated by both vertebrate and invertebrate β subunits (Jeziorski et al., 1999). Furthermore, w10c8.1 cannot block the modulatory effects of exogenous β subunits, suggesting that the C. elegans protein does not compete for the β subunit interaction site on the α1 subunit (M. C. Jeziorski, unpublished data). The robust effects exerted upon the Cyanea α1 subunit by other β subunits occurs in spite of sequence variation within the interaction domain of the α1 subunit, however, and the w10c8.1 protein may bind more efficiently to an interaction domain that is more highly conserved or to a domain specific to a C. elegans HVA α1 subunit. Alternatively, the β homologue may modulate other channel proteins found in C. elegans, such as the T-type channel or the more distant α1 subunit homologues of unknown function.

The C. elegans genome also contains two genes encoding putative α2δ subunits. One (accession number P34374) corresponds to the unc-36 gene, mutation of which results in a phenotype similar to that for unc-2 (Schafer et al., 1996). The published sequence of 734 amino acids, which is considerably shorter than the average length of mammalian α2δ subunits (approximately 1100 residues) and lacks the hydrophobic δ segment that anchors the α2 segment to the membrane, is almost certainly incomplete. The second identified gene (accession number CAA90091) encodes what appears to be a full-length protein. A phylogenetic analysis of known α2δ subunits (truncated to match the length of the unc-36 gene product) indicates that the two C. elegans genes are approximately as distant from each other as they are from mammalian isoforms.

Although the genetic information from C. elegans is far more extensive than that for any other multicellular organism, it may not accurately reflect the full complement of Ca2+ channel subunits present in all invertebrates. No copy of a pore-forming α subunit gene for a voltage-gated Na+ channel is found in C. elegans, for instance. The absence of such a gene is probably due to its deletion during the evolution of nematodes, because Na+ channels have been cloned from earlier organisms such as cnidarians (Anderson et al., 1993; Spafford et al., 1998; White et al., 1998b) and platyhelminths (Jeziorski et al., 1997) as well as from higher invertebrates such as Drosophila (Salkoff et al., 1987) and squid (Sato and Matsumoto, 1992; Rosenthal and Gilly, 1993). A similar deletion may have occurred for one or more Ca2+ channel subunit genes, such as a gene encoding a γ subunit, which is not represented in the C. elegans genome.

In this regard, the report of two additional cDNA fragments amplified by the polymerase chain reaction (PCR) from the coral Stylophora pistillata (Zoccola et al., 1999) deserves mention. The fragments, each approximately 100 codons in size, appear to encode the domain IV pore and S6 regions of two different non-L-type α1 subunits. The two sequences have been labeled as N-type and P-type, suggesting that divergence within the non-L-type channel subfamily occurred very early in metazoan evolution. Although the limited size of the fragments precludes their definitive identification, our preliminary phylogenetic analysis places the two sequences on adjacent branches among invertebrate non-L-type α1 subunits, and not within the mammalian α1A or α1B clusters (data not shown). Nevertheless, the potential existence of multiple non-L-type α1 subunit genes in an early invertebrate demonstrates that current models of channel evolution may require modification and emphasizes the importance of the continued isolation and characterization of invertebrate channel subunit genes.

The C. elegans genome clearly lacks the variety of Ca2+ channel subunit genes that underlies much of the heterogeneity in vertebrate Ca2+ currents. The multiple mammalian isoforms of L-type, non-L-type and T-type α1 subunits are each represented by single genes in C. elegans, and perhaps in most invertebrates. Numerous potential combinations between the four vertebrate β isoforms and the HVA α1 subunits provide additional channel variation that cannot be mimicked by a single invertebrate β subunit. How might multiple Ca2+ currents be generated from fewer subunit genes?

One possible source of distinct currents is through post-translational modification. Most invertebrate subunit sequences contain consensus sites for phosphorylation by protein kinases A and C, among other kinases, and phosphorylation has been extensively implicated in the modulation of invertebrate Ca2+ current amplitude (for a review, see Kits and Mansvelder, 1996). Stimulation of protein kinase C also causes the non-L-type Ca2+ channel in Aplysia bag cell neurons to translocate from vesicular to plasma membranes (White et al., 1998a). The ability of phosphorylation to generate multiple current components that differ in their activation or inactivation kinetics or pharmacological response has not been extensively investigated, however.

In contrast, alternative splicing of subunit transcripts, which is extensively employed in mammalian tissues to generate variant Ca2+ channels (for a review, see Perez-Reyes and Schneider, 1995), has been carefully examined in Drosophila (Smith et al., 1996; Peixoto et al., 1997). The non-L-type α1 subunit gene, known as Dmca1A, varies at three key sites (Peixoto et al., 1997). Two involve the S4 segments critical to determining the voltage-dependence of the channel. The loop connecting IS3 and IS4, as well as much of IS4 itself, is encoded in two separate exons, while the IVS3–IVS4 region contains a six-base insertion in some transcripts. Variation at each site may alter channel activation. Another intriguing splice site occurs within the intracellular loop between domains I and II. A portion of this region that is highly conserved in vertebrate α1 subunits forms both the primary interaction site for the β subunit and the site to which βγ subunits of G-proteins bind in certain non-L-type channels. Whereas one exon in Dmca1A contains the fully conserved interaction sites, a second exon of identical length is lacking the motif to which Gβγ subunits bind and is conserved with the canonical β subunit binding domain at only four of nine residues. Three of the remaining residues are those most critical to the α1–β interaction (De Waard et al., 1996), and the Cyanea α1 subunit, which also contains a poorly conserved I–II loop with the key residues intact, can be modulated by β subunits (Jeziorski et al., 1999). Nevertheless, variation within this interaction domain may result in subtle influences upon channel kinetics or membrane targeting. Although it has been suggested that the two exons in the I–II loop arose early in the evolution of non-L-type channels (Peixoto et al., 1997), the introns flanking the corresponding exon in the Cyanea L-type channel gene do not appear to contain an alternative exon for the β binding region (M. C. Jeziorski, R. M. Greenberg and P. A. V. Anderson, unpublished observations). Invertebrate β subunits may also be alternatively spliced, as demonstrated in the Musca gene (Grabner et al., 1994b).

An additional way in which multiple channels may be derived from one invertebrate gene is via RNA editing. In certain genes, specific adenosine residues in the genomic sequence can be converted to guanosines in the unprocessed mRNA, resulting in altered translation at discrete sites (for a review, see Simpson and Emeson, 1996). RNA editing has been extensively implicated in the regulation of mammalian glutamate channels (Sommer et al., 1991; Lomeli et al., 1994) and, in squid, functionally distinct K+ channels are generated by editing of transcripts from a single gene (Patton et al., 1997). RNA editing may also underlie heterogeneity of invertebrate Ca2+ currents. Transcripts from the Drosophila Dmca1A gene can be edited at ten different sites, resulting in an amino acid substitution at each position (Smith et al., 1996; Peixoto et al., 1997). The proportion of edited compared with unedited codons at each site is not yet known, but many of the affected residues lie within functionally critical regions such as transmembrane segments or pore-forming loops. The potential for differential editing of specific residues in different cell types or in response to developmental or environmental cues could result in significant channel diversity from a limited number of genes.

The path by which Ca2+ channels evolved is implied by their relationship with other members of the voltage-gated cation channel superfamily. Each domain of the pore-forming subunits of Ca2+ and Na+ channels is similar in structure to a K+ channel subunit, suggesting that the form of the K+ channel containing six transmembrane segments evolved first, then underwent two stages of concatenation to create an early four-domain protein. One proposed reason for the transition from single-domain to four-domain proteins is that an asymmetrical pore structure may be required in Ca2+ and Na+ channels (Salkoff and Jegla, 1995). The four-domain protein may have given rise to Ca2+-selective channels before Na+ channels. This pattern of emerging channel selectivity is thought to correspond to the demands imposed upon organisms as they gained greater complexity (Hille, 1992; Ranganathan, 1994; Armstrong and Hille, 1998). Early cells may have employed K+-selective channels as a defense against osmotic stress (Milkman, 1994; Sukharev et al., 1994). The need to control intracellular Ca2+ concentration perhaps resulted from the ability of free Ca2+ to precipitate phosphate and to disrupt the cell’s primary source of metabolic energy; the role of Ca2+ as a second messenger, and the subsequent need for gated Ca2+ influx, may have developed from such regulatory mechanisms (Clapham, 1995). The requirement for rapid signal propagation in the earliest nervous systems would then have necessitated a role for Na+-permeable channels.

Current physiological and molecular evidence supports several elements of this evolutionary model. The structure of the single domain apparently developed before prokaryotes and eukaryotes diverged; the kch protein cloned from Escherichia coli possesses the six transmembrane segments and conserved pore region characteristic of eukaryotic K+ channels (Milkman, 1994). Subsequent within-gene duplication may have occurred prior to the separation of metazoans from other eukaryotes, since the four-domain structure is evident in the two yeast channels described above. Conceivably, either the single-domain or four-domain structures could have evolved independently in two separate lineages. However, homology among voltage-gated channels has typically been inferred from sequence identity, which is less likely to arise by convergent evolution than is structural similarity. The ionic selectivity of the early four-domain channels is unknown, but it has been proposed that the yeast Cch1 protein gates Ca2+ (Paidhungat and Garrett, 1997), and the structure of its pore is somewhat consistent with Ca2+ selectivity. Other physiological evidence also indirectly suggests that Ca2+ currents predate Na+ currents. The existence of Ca2+ currents in protists such as Paramecium (for a review, see Kung, 1989) and Stylonychia mytilus (Deitmer, 1986) has been firmly established, and Ca2+ currents have been described in plant cells as well (Thuleau et al., 1994). The earliest phylum in which voltage-gated Na+ currents have been extensively recorded is Cnidaria (Mackie and Meech, 1985; Anderson, 1987), although the report of a protist action potential that is carried in part by Na+ (Febvre-Chevalier et al., 1989) indicates that Na+-selective channels may have evolved prior to metazoans. Sequence similarity between Na+ and Ca2+ channels, which is greater than that between either class and the yeast or Rb21 proteins, argues for a relatively recent division of the two families. Phylogenetic analysis suggests that the ancestral four-domain channel gave rise first to the Rb21 family, then to a precursor of metazoan Ca2+ and Na+ channels. Na+ channels evolved either from this branch or from the LVA Ca2+ channel subfamily after it separated from HVA channels. Definitive support for this proposed series of evolutionary divergences awaits isolation of channel sequences from protists and other early eukaryotes. Furthermore, we commonly infer the type of current gated by a channel from the overall sequence identity of the protein, but minimal substitution within the pore is required to alter the selectivity of a channel. Thus, a channel subfamily associated with a particular current may have evolved initially to gate a different ion or mixture of ions before assuming its present role.

From the earliest Ca2+ channel, two additional phases of divergence occurred, one to generate the three (or more) major Ca2+ channel types, and a second to create the various subtypes found in mammals. Similarly, present models of metazoan evolution propose two major stages of genome expansion. The first, which occurred prior to the separation of parazoans (sponges) and true metazoans, permitted the development of functionally distinct isoforms of a given protein, thus facilitating differentiation of tissues within an organism. Sponges, which lack discernible tissues, nevertheless possess multiple genes for proteins such as protein kinase C, phospholipase C (Koyanagi et al., 1998), protein tyrosine kinases (Suga et al., 1999) and protein kinase phosphatases (Ono et al., 1999), isoforms that correspond to gene subfamilies found in higher organisms. The second genomic expansion occurred early in chordate evolution, when two successive genome duplications freed gene subtypes to diversify further, leading to the additional genetic variation that predated the appearance of vertebrates (Holland et al., 1994). The protein subtypes that resulted from the latter duplication event typically differ less from each other than do the subfamilies from which they emerged, and they are thought to vary more in tissue distribution than in function (Iwabe et al., 1996).

Evidence from invertebrate Ca2+ channel genes is consistent with the timing of the two duplication events. At least three classes of Ca2+ channel α1 subunit genes (L, non-L and T-type) are represented in C. elegans, and the α1 subunits cloned from cnidarians, the earliest true metazoans, fall within specific channel subfamilies, indicating that the appearance of multiple α1 subunit genes predates this phylum. A similar organization is suggested at the opposite end of the invertebrate spectrum, where the α1 subunit cloned from the invertebrate chordate Halocynthia can be identified as L-type, but not as a specific subtype. The invertebrate β subunits that have been cloned also appeared prior to the divergence of mammalian subtypes (Grabner et al., 1994b; Jeziorski et al., 1999). The earliest clear evidence for subtypes within the larger classes of subunit genes is found in nonmammalian vertebrates. Neurons of the marine ray Discopyge ommata express at least two non-L-type subtypes, α1B and α1E (Horne et al., 1993), and a β subunit cloned from Xenopus corresponds to the β3 mammalian subtype (Tareilus et al., 1997). Comparison of invertebrate and vertebrate genes supports the model of a fourfold increase in diversity prior to the evolution of vertebrates. The L-, non-L- and T-type α1 subunit genes in the C. elegans genome are represented, respectively, by four, three and three known mammalian subtypes. Similarly, the four vertebrate β subunit subtypes exhibit comparable identity to the lone conventional C. elegans β subunit gene; clustering of the vertebrate subunits into two pairs may reflect the two rounds of genome duplication that have been proposed.

The gene duplication model cannot be expected to be followed strictly by all gene families. Several potential inconsistencies are already apparent, such as the absence of C. elegans genes encoding a Na+ channel α subunit or a Ca2+ channel γ subunit, and the existence of two α2δ subunit homologues in C. elegans, each equally distant from known mammalian proteins. Nevertheless, the recent discovery of a fourth mammalian L-type α1 subunit suggests that additional non-L- or T-type channel genes may eventually be found and that the novel putative channel subunits found in C. elegans may also be revealed to have multiple vertebrate homologues.

It is clear from the advances described in this review that a survey of Ca2+ channel subtypes can no longer be complete without consideration of invertebrate channels. Information from the various nonmammalian genome projects may offer the most immediate and tangible benefits, but much can be learned from channels in other organisms as well. Invertebrate genes separated several hundred million years ago from the ancestral genes that gave rise to mammalian channels and, since that time, have evolved independently, within the constraints of maintaining a functional Ca2+ channel. The resulting variance between vertebrate and invertebrate peptide sequences more clearly defines the conserved regions essential to channel function while revealing structural variation that may underlie distinct physiological characteristics. When combined with the ability to analyze and manipulate genetic information rapidly in established model systems, the growing understanding of the molecular biology of invertebrate Ca2+ channels promises to lead to an integrated view of Ca2+ channel function and diversity.

We thank Dr Peter Okkema and Dr William Walthall for supplying C. elegans cDNA libraries, Dr Edward Perez-Reyes for contributing mammalian Ca2+ channel subunit cDNA clones and Dr Carlos Valverde and Dr Aurea Orozco for comments on the manuscript. This work was supported in part by NSF grant IBN-9808386 to P.A.V.A. and R.M.G.

Altschul
,
S. F.
,
Gish
,
W.
,
Miller
,
W.
,
Myers
,
E. W.
and
Lipman
,
D. J.
(
1990
).
Basic local alignment search tool
.
J. Mol. Biol.
215
,
403
410
.
Anderson
,
P. A. V.
(
1987
).
Properties and pharmacology of a TTX-insensitive Na+ current in neurones of the jellyfish Cyanea capillata
.
J. Exp. Biol.
133
,
231
248
.
Anderson
,
P. A. V.
,
Holman
,
M. A.
and
Greenberg
,
R. M.
(
1993
).
Deduced amino acid sequence of a putative sodium channel from the scyphozoan jellyfish Cyanea capillata
.
Proc. Natl. Acad. Sci. USA
90
,
7419
7423
.
Araque
,
A.
,
Clarac
,
F.
and
Buno
,
W.
(
1994
).
P-type Ca2+ channels mediate excitatory and inhibitory synaptic transmitter release in crayfish muscle
.
Proc. Natl. Acad. Sci. USA
91
,
4224
4228
.
Armstrong
,
C. M.
and
Hille
,
B.
(
1998
).
Voltage-gated ion channels and electrical excitability
.
Neuron
20
,
371
380
.
Arnoult
,
C.
and
Villaz
,
M.
(
1994
).
Differential developmental fates of the two calcium currents in early embryos of the ascidian Ciona intestinalis
.
J. Membr. Biol.
137
,
127
135
.
Bargmann
,
C. I.
(
1998
).
Neurobiology of the Caenorhabditis elegans genome
.
Science
282
,
2028
2033
.
Bech-Hansen
,
N. T.
,
Naylor
,
M. J.
,
Maybaum
,
T. A.
,
Pearce
,
W. G.
,
Koop
,
B.
,
Fishman
,
G. A.
,
Mets
,
M.
,
Musarella
,
M. A.
and
Boycott
,
K. M.
(
1998
).
Loss-of-function mutations in a calcium-channel α1-subunit gene in Xp11.23 cause incomplete X-linked congenital stationary night blindness
.
Nature Genetics
19
,
264
267
.
Bickmeyer
,
U.
,
Rossler
,
W.
and
Wiegand
,
H.
(
1994
).
ω AGA toxin IVA blocks high-voltage-activated calcium channel currents in cultured pars intercerebralis neurosecretory cells of adult Locusta migratoria
.
Neurosci. Lett.
181
,
113
116
.
Brezina
,
V.
,
Evans
,
C. G.
and
Weiss
,
K. R.
(
1994
).
Characterization of the membrane ion currents of a model molluscan muscle, the accessory radula closer muscle of Aplysia californica. III. Depolarization-activated Ca current
.
J. Neurophysiol.
71
,
2126
2138
.
Castellano
,
A.
,
Wei
,
X.
,
Birnbaumer
,
L.
and
Perez-Reyes
,
E.
(
1993a
).
Cloning and expression of a third calcium channel β subunit
.
J. Biol. Chem.
268
,
3450
3455
.
Castellano
,
A.
,
Wei
,
X.
,
Birnbaumer
,
L.
and
Perez-Reyes
,
E.
(
1993b
).
Cloning and expression of a neuronal calcium channel β subunit
.
J. Biol. Chem.
268
,
12359
12366
.
Catterall
,
W. A.
(
1995
).
Structure and function of voltage-gated ion channels
.
Annu. Rev. Biochem.
64
,
493
531
.
Chien
,
A. J.
,
Zhao
,
X.
,
Shirokov
,
R. E.
,
Puri
,
T. S.
,
Chang
,
C. F.
,
Sun
,
D.
,
Rios
,
E.
and
Hosey
,
M. M.
(
1995
).
Roles of a membrane-localized β subunit in the formation and targeting of functional L-type Ca2+ channels
.
J. Biol. Chem.
270
,
30036
30044
.
Clapham
,
D. E.
(
1995
).
Calcium signaling
.
Cell
80
,
259
268
.
Cribbs
,
L. L.
,
Lee
,
J. H.
,
Yang
,
J.
,
Satin
,
J.
,
Zhang
,
Y.
,
Daud
,
A.
,
Barclay
,
J.
,
Williamson
,
M. P.
,
Fox
,
M.
,
Rees
,
M.
and
Perez-Reyes
,
E.
(
1998
).
Cloning and characterization of α1H from human heart, a member of the T-type Ca2+ channel gene family
.
Circ. Res.
83
,
103
109
.
Deitmer
,
J. W.
(
1986
).
Voltage dependence of two inward currents carried by calcium and barium in the ciliate Stylonychia mytilus
.
J. Physiol., Lond.
380
,
551
574
.
De Waard
,
M.
,
Liu
,
H.
,
Walker
,
D.
,
Scott
,
V. E.
,
Gurnett
,
C. A.
and
Campbell
,
K. P.
(
1997
).
Direct binding of G-protein βγ complex to voltage-dependent calcium channels
.
Nature
385
,
446
450
.
De Waard
,
M.
,
Scott
,
V. E. S.
,
Pragnell
,
M.
and
Campbell
,
K. P.
(
1996
).
Identification of critical amino acids involved in α1–β interaction in voltage-dependent Ca2+ channels
.
FEBS Lett.
380
,
272
276
.
Dolmetsch
,
R. E.
,
Lewis
,
R. S.
,
Goodnow
,
C. C.
and
Healy
,
J. I.
(
1997
).
Differential activation of transcription factors induced by Ca2+ response amplitude and duration
.
Nature
386
,
855
858
.
Dreijer
,
A. M.
and
Kits
,
K. S.
(
1995
).
Multiple second messenger routes enhance two high-voltage-activated calcium currents in molluscan neuroendocrine cells
.
Neuroscience
64
,
787
800
.
Dubel
,
S. J.
,
Starr
,
T. V.
,
Hell
,
J.
,
Ahlijanian
,
M. K.
,
Enyeart
,
J. J.
,
Catterall
,
W. A.
and
Snutch
,
T. P.
(
1992
).
Molecular cloning of the α1 subunit of an ω-conotoxin-sensitive calcium channel
.
Proc. Natl. Acad. Sci. USA
89
,
5058
5062
.
Eberl
,
D. F.
,
Ren
,
D.
,
Feng
,
G.
,
Lorenz
,
L. J.
,
Van Vactor
,
D.
and
Hall
,
L. M.
(
1998
).
Genetic and developmental characterization of Dmca1D, a calcium channel α1 subunit gene in Drosophila melanogaster
.
Genetics
148
,
1159
1169
.
Ellinor
,
P. T.
,
Yang
,
J.
,
Sather
,
W. A.
,
Zhang
,
J. F.
and
Tsien
,
R. W.
(
1995
).
Ca2+ channel selectivity at a single locus for highaffinity Ca2+ interactions
.
Neuron
15
,
1121
1132
.
Ellinor
,
P. T.
,
Zhang
,
J. F.
,
Horne
,
W. A.
and
Tsien
,
R. W.
(
1994
).
Structural determinants of the blockade of N-type calcium channels by a peptide neurotoxin
.
Nature
372
,
272
275
.
Ellis
,
S. B.
,
Williams
,
M. E.
,
Ways
,
N. R.
,
Brenner
,
R.
,
Sharp
,
A. H.
,
Leung
,
A. T.
,
Campbell
,
K. P.
,
McKenna
,
E.
,
Koch
,
W. J.
,
Hui
,
A.
,
Schwartz
,
A.
and
Harpold
,
M. M.
(
1988
).
Sequence and expression of mRNAs encoding the α1 and α2 subunits of a DHP-sensitive calcium channel
.
Science
241
,
1661
1664
.
Erxleben
,
C.
and
Rathmayer
,
W.
(
1997
).
A dihydropyridinesensitive voltage-dependent calcium channel in the sarcolemmal membrane of crustacean muscle
.
J. Gen. Physiol.
109
,
313
326
.
Fatt
,
P.
and
Ginsborg
,
B. L.
(
1958
).
The ionic requirement for the production of action potentials in crustacean muscle fibres
.
J. Physiol., Lond.
142
,
516
543
.
Febvre-Chevalier
,
C.
,
Bilbaut
,
A.
,
Febvre
,
J.
and
Bone
,
Q.
(
1989
).
Membrane excitability and motile responses in the Protozoa, with particular attention to the heliozoan Actinocoryne contractilis
. In
Evolution of the First Nervous Systems
(ed.
P. A. V.
Anderson
), pp.
237
253
. New York: Plenum Press.
Fieber
,
L. A.
(
1998
).
Characterization of Na+ and Ca2+ currents in bag cells of sexually immature Aplysia californica
.
J. Exp. Biol.
201
,
745
754
.
Fossier
,
P.
,
Baux
,
G.
and
Tauc
,
L.
(
1994
).
N- and P-type Ca2+ channels are involved in acetylcholine release at a neuroneuronal synapse: only the N-type channel is the target of neuromodulators
.
Proc. Natl. Acad. Sci. USA
91
,
4771
4775
.
García-Colunga
,
J.
,
Valdiosera
,
R.
and
García
,
U.
(
1999
).
P-type Ca2+ current in crayfish peptidergic neurones
.
J. Exp. Biol.
202
,
429
440
.
Gielow
,
M. L.
,
Gu
,
G. G.
and
Singh
,
S.
(
1995
).
Resolution and pharmacological analysis of the voltage-dependent calcium channels of Drosophila larval muscles
.
J. Neurosci.
15
,
6085
6093
.
Grabner
,
M.
,
Bachmann
,
A.
,
Rosenthal
,
F.
,
Striessnig
,
J.
,
Schultz
,
C.
,
Tautz
,
D.
and
Glossmann
,
H.
(
1994a
).
Molecular cloning of an α1-subunit from housefly (Musca domestica) muscle
.
FEBS Lett.
339
,
189
194
.
Grabner
,
M.
,
Dirksen
,
R. T.
,
Suda
,
N.
and
Beam
,
K. G.
(
1999
).
The II–III loop of the skeletal muscle dihydropyridine receptor is responsible for the bi-directional coupling with the ryanodine receptor
.
J. Biol. Chem.
274
,
21913
21919
.
Grabner
,
M.
,
Wang
,
Z.
,
Hering
,
S.
,
Striessnig
,
J.
and
Glossmann
,
H.
(
1996
).
Transfer of 1,4-dihydropyridine sensitivity from L-type to class A (BI) calcium channels
.
Neuron
16
,
207
218
.
Grabner
,
M.
,
Wang
,
Z.
,
Mitterdorfer
,
J.
,
Rosenthal
,
F.
,
Charnet
,
P.
,
Savchenko
,
A.
,
Hering
,
S.
,
Ren
,
D.
,
Hall
,
L. M.
and
Glossmann
,
H.
(
1994b
).
Cloning and functional expression of a neuronal calcium channel β subunit from house fly (Musca domestica)
.
J. Biol. Chem.
269
,
23668
23674
.
Greenberg
,
R. M.
,
Clark
,
K. S.
,
Jeziorski
,
M. C.
,
White
,
G. B.
and
Anderson
,
P. A. V.
(
1995
).
Structure of calcium channel α1 subunits from cnidarians and platyhelminths (abstract)
.
Soc. Neurosci. Abstr.
21
,
1371
.
Grolleau
,
F.
and
Lapied
,
B.
(
1996
).
Two distinct low-voltageactivated Ca2+ currents contribute to the pacemaker mechanism in cockroach dorsal unpaired median neurons
.
J. Neurophysiol.
76
,
963
976
.
Gu
,
X.
and
Spitzer
,
N. C.
(
1995
).
Distinct aspects of neuronal differentiation encoded by frequency of spontaneous Ca2+ transients
.
Nature
375
,
784
787
.
Hagiwara
,
S.
,
Ozawa
,
S.
and
Sand
,
O.
(
1975
).
Voltage clamp analysis of two inward current mechanisms in the egg cell membrane of a starfish
.
J. Gen. Physiol.
65
,
617
644
.
Haydon
,
P. G.
and
Man-Son-Hing
,
H.
(
1988
).
Low- and highvoltage-activated calcium currents: their relationship to the site of neurotransmitter release in an identified neuron of Helisoma
.
Neuron
1
,
919
927
.
Heinemann
,
S. H.
,
Terlau
,
H.
,
Stuhmer
,
W.
,
Imoto
,
K.
and
Numa
,
S.
(
1992
).
Calcium channel characteristics conferred on the sodium channel by single mutations
.
Nature
356
,
441
443
.
Hell
,
J. W.
,
Westenbroek
,
R. E.
,
Warner
,
C.
,
Ahlijanian
,
M. K.
,
Prystay
,
W.
,
Gilbert
,
M. M.
,
Snutch
,
T. P.
and
Catterall
,
W. A.
(
1993
).
Identification and differential subcellular localization of the neuronal class C and class D L-type calcium channel α1 subunits
.
J. Cell Biol.
123
,
949
962
.
Hille
,
B.
(
1992
).
Ionic Channels of Excitable Membranes, 2nd edition. Sunderland, MA: Sinauer Assoc
.
Hockerman
,
G. H.
,
Johnson
,
B. D.
,
Scheuer
,
T.
and
Catterall
,
W. A.
(
1995
).
Molecular determinants of high affinity phenylalkylamine block of L-type calcium channels
.
J. Biol. Chem.
270
,
22119
22122
.
Hockerman
,
G. H.
,
Peterson
,
B. Z.
,
Johnson
,
B. D.
and
Catterall
,
W. A.
(
1997
).
Molecular determinants of drug binding and action on L-type calcium channels
.
Annu. Rev. Pharmac. Toxicol.
37
,
361
396
.
Holland
,
P. W. H.
,
Garcia-Fernandez
,
J.
,
Williams
,
N. A.
and
Sidow
,
A.
(
1994
).
Gene duplications and the origins of vertebrate development
.
Development (Suppl.), 125–133
.
Hong
,
S. J.
and
Lnenicka
,
G. A.
(
1997
).
Characterization of a P-type calcium current in a crayfish motoneuron and its selective modulation by impulse activity
.
J. Neurophysiol.
77
,
76
85
.
Horne
,
W. A.
,
Ellinor
,
P. T.
,
Inman
,
I.
,
Zhou
,
M.
,
Tsien
,
R. W.
and
Schwarz
,
T. L.
(
1993
).
Molecular diversity of Ca2+ channel α1 subunits from the marine ray Discopyge ommata
.
Proc. Natl. Acad. Sci. USA
90
,
3787
3791
.
Hui
,
A.
,
Ellinor
,
P. T.
,
Krizanova
,
O.
,
Wang
,
J. J.
,
Diebold
,
R. J.
and
Schwartz
,
A.
(
1991
).
Molecular cloning of multiple subtypes of a novel rat brain isoform of the α1 subunit of the voltage-dependent calcium channel
.
Neuron
7
,
35
44
.
Inagaki
,
S.
,
Kaku
,
K.
,
Dunlap
,
D. Y.
and
Matsumura
,
F.
(
1998
).
Sequences of cDNAs encoding calmodulin and partial structures of calmodulin kinase and a calcium channel of kdr-resistant and -susceptible German cockroaches, Blattella germanica
.
Comp. Biochem. Physiol.
120C
,
225
233
.
Iwabe
,
N.
,
Kuma
,
K.
and
Miyata
,
T.
(
1996
).
Evolution of gene families and relationship with organismal evolution: rapid divergence of tissue-specific genes in the early evolution of chordates
.
Mol. Biol. Evol.
13
,
483
493
.
Jay
,
S. D.
,
Ellis
,
S. B.
,
McCue
,
A. F.
,
Williams
,
M. E.
,
Vedvick
,
T. S.
,
Harpold
,
M. M.
and
Campbell
,
K. P.
(
1990
).
Primary structure of the γ subunit of the DHP-sensitive calcium channel from skeletal muscle
.
Science
248
,
490
492
.
Jeziorski
,
M. C.
,
Greenberg
,
R. M.
and
Anderson
,
P. A.
(
1997
).
Cloning of a putative voltage-gated sodium channel from the turbellarian flatworm Bdelloura candida
.
Parasitology
115
,
289
296
.
Jeziorski
,
M. C.
,
Greenberg
,
R. M.
and
Anderson
,
P. A. V.
(
1999
).
Cloning and expression of a jellyfish calcium channel β subunit reveal functional conservation of the α1–β interaction
.
Receptors and Channels
6
,
375
386
.
Jeziorski
,
M. C.
,
Greenberg
,
R. M.
,
Clark
,
K. S.
and
Anderson
,
P. A.
(
1998
).
Cloning and functional expression of a voltage-gated calcium channel α1 subunit from jellyfish
.
J. Biol. Chem.
273
,
22792
22799
.
Jones
,
S. W.
(
1998
).
Overview of voltage-dependent calcium channels
.
J. Bioenerg. Biomembr.
30
,
299
312
.
Kimura
,
T.
,
Shouno
,
O.
,
Hirota
,
K.
,
Saito
,
T.
,
Matsumoto
,
G.
and
Sato
,
C.
(
1997
).
Molecular cloning and characterization of a putative neural calcium channel α1-subunit from squid optic lobe
.
Biochem. Biophys. Res. Commun.
230
,
147
154
.
Kits
,
K. S.
and
Mansvelder
,
H. D.
(
1996
).
Voltage gated calcium channels in molluscs: classification, Ca2+ dependent inactivation, modulation and functional roles
.
Invert. Neurosci.
2
,
9
34
.
Klugbauer
,
N.
,
Lacinová
,
L.
,
Marais
,
E.
,
Hobom
,
M.
and
Hofmann
,
F.
(
1999
).
Molecular diversity of the calcium channel α2δ subunit
.
J. Neurosci.
19
,
684
691
.
Koyanagi
,
M.
,
Ono
,
K.
,
Suga
,
H.
,
Iwabe
,
N.
and
Miyata
,
T.
(
1998
).
Phospholipase C cDNAs from sponge and hydra: antiquity of genes involved in the inositol phospholipid signaling pathway
.
FEBS Lett.
439
,
66
70
.
Kraus
,
R. L.
,
Hering
,
S.
,
Grabner
,
M.
,
Ostler
,
D.
and
Striessnig
,
J.
(
1998
).
Molecular mechanism of diltiazem interaction with L-type Ca2+ channels
.
J. Biol. Chem.
273
,
27205
27212
.
Kung
,
C.
(
1989
).
Ion channels of unicellular microbes
. In
Evolution of the First Nervous Systems
(ed.
P. A. V.
Anderson
), pp.
203
235
.
New York
:
Plenum Press
.
Lacerda
,
A. E.
,
Kim
,
H. S.
,
Ruth
,
P.
,
Perez-Reyes
,
E.
,
Flockerzi
,
V.
,
Hofmann
,
F.
,
Birnbaumer
,
L.
and
Brown
,
A. M.
(
1991
).
Normalization of current kinetics by interaction between the α1 and β subunits of the skeletal muscle dihydropyridine-sensitive Ca2+ channel
.
Nature
352
,
527
530
.
Laurienti
,
P. J.
and
Blankenship
,
J. E.
(
1997
).
Serotonergic modulation of a voltage-gated calcium current in parapodial swim muscle from Aplysia brasiliana
.
J. Neurophysiol.
77
,
1496
1502
.
Lee
,
J. H.
,
Cribbs
,
L. L.
and
Perez-Reyes
,
E.
(
1999a
).
Cloning of a novel four repeat protein related to voltage-gated sodium and calcium channels
.
FEBS Lett.
445
,
231
236
.
Lee
,
J. H.
,
Daud
,
A. N.
,
Cribbs
,
L. L.
,
Lacerda
,
A. E.
,
Pereverzev
,
A.
,
Klockner
,
U.
,
Schneider
,
T.
and
Perez-Reyes
,
E.
(
1999b
).
Cloning and expression of a novel member of the low voltage-activated T-type calcium channel family
.
J. Neurosci.
19
,
1912
1921
.
Lee
,
R. Y. N.
,
Lobel
,
L.
,
Hengartner
,
M.
,
Horvitz
,
H. R.
and
Avery
,
L.
(
1997
).
Mutations in the α1 subunit of an L-type voltage-activated Ca2+ channel cause myotonia in Caenorhabditis elegans
.
EMBO J.
16
,
6066
6076
.
Letts
,
V. A.
,
Felix
,
R.
,
Biddlecome
,
G. H.
,
Arikkath
,
J.
,
Mahaffey
,
C. L.
,
Valenzuela
,
A.
,
Bartlett II
,
F. S.
,
Mori
,
Y.
,
Campbell
,
K. P.
and
Frankel
,
W. N.
(
1998
).
The mouse stargazer gene encodes a neuronal Ca2+-channel γ subunit
.
Nature Genetics
19
,
340
347
.
Leung
,
H. T.
,
Branton
,
W. D.
,
Phillips
,
H. S.
,
Jan
,
L.
and
Byerly
,
L.
(
1989
).
Spider toxins selectively block calcium currents in Drosophila
.
Neuron
3
,
767
772
.
Liu
,
H.
,
De Waard
,
M.
,
Scott
,
V. E. S.
,
Gurnett
,
C. A.
,
Lennon
,
V. A.
and
Campbell
,
K. P.
(
1996
).
Identification of three subunits of the high affinity ω-conotoxin MVIIC-sensitive Ca2+ channel
.
J. Biol. Chem.
271
,
13804
13810
.
Llinas
,
R.
,
Sugimori
,
M.
,
Lin
,
J. W.
and
Cherksey
,
B.
(
1989
).
Blocking and isolation of a calcium channel from neurons in mammals and cephalopods utilizing a toxin fraction (FTX) from funnel-web spider poison
.
Proc. Natl. Acad. Sci. USA
86
,
1689
1693
.
Lomeli
,
H.
,
Mosbacher
,
J.
,
Melcher
,
T.
,
Hoger
,
T.
,
Geiger
,
J. R.
,
Kuner
,
T.
,
Monyer
,
H.
,
Higuchi
,
M.
,
Bach
,
A.
and
Seeburg
,
P. H.
(
1994
).
Control of kinetic properties of AMPA receptor channels by nuclear RNA editing
.
Science
266
,
1709
1713
.
Mackie
,
G. O.
and
Meech
,
R. W.
(
1985
).
Separate sodium and calcium spikes in the same axon
.
Nature
313
,
791
793
.
McFarlane
,
M. B.
and
Gilly
,
W. F.
(
1996
).
Spatial localization of calcium channels in giant fiber lobe neurons of the squid (Loligo opalescens)
.
Proc. Natl. Acad. Sci. USA
93
,
5067
5071
.
Mikami
,
A.
,
Imoto
,
K.
,
Tanabe
,
T.
,
Niidome
,
T.
,
Mori
,
Y.
,
Takeshima
,
H.
,
Narumiya
,
S.
and
Numa
,
S.
(
1989
).
Primary structure and functional expression of the cardiac dihydropyridinesensitive calcium channel
.
Nature
340
,
230
233
.
Milkman
,
R.
(
1994
).
An Escherichia coli homologue of eukaryotic potassium channel proteins
.
Proc. Natl. Acad. Sci. USA
91
,
3510
3514
.
Mills
,
J. D.
and
Pitman
,
R. M.
(
1997
).
Electrical properties of a cockroach motor neuron soma depend on different characteristics of individual Ca components
.
J. Neurophysiol.
78
,
2455
2466
.
Mitterdorfer
,
J.
,
Wang
,
Z.
,
Sinnegger
,
M. J.
,
Hering
,
S.
,
Striessnig
,
J.
,
Grabner
,
M.
and
Glossmann
,
H.
(
1996
).
Two amino acid residues in the IIIS5 segment of L-type calcium channels differentially contribute to 1,4-dihydropyridine sensitivity
.
J. Biol. Chem.
271
,
30330
30335
.
Morales
,
M.
,
Ferrus
,
A.
and
Martinez-Padron
,
M.
(
1999
).
Presynaptic calcium-channel currents in normal and csp mutant Drosophila peptidergic terminals
.
Eur. J. Neurosci.
11
,
1818
1826
.
Mori
,
Y.
,
Friedrich
,
T.
,
Kim
,
M.-S.
,
Mikami
,
A.
,
Nakai
,
J.
,
Ruth
,
P.
,
Bosse
,
E.
,
Hofmann
,
F.
,
Flockerzi
,
V.
,
Furuichi
,
T.
,
Mikoshiba
,
K.
,
Imoto
,
K.
,
Tanabe
,
T.
and
Numa
,
S.
(
1991
).
Primary structure and functional expression from complementary DNA of a brain calcium channel
.
Nature
350
,
398
402
.
Nakai
,
J.
,
Tanabe
,
T.
,
Konno
,
T.
,
Adams
,
B.
and
Beam
,
K. G.
(
1998
).
Localization in the II–III loop of the dihydropyridine receptor of a sequence critical for excitation–contraction coupling
.
J. Biol. Chem.
273
,
24983
24986
.
Nerbonne
,
J. M.
and
Gurney
,
A. M.
(
1987
).
Blockade of Ca2+ and K+ currents in bag cell neurons of Aplysia californica by dihydropyridine Ca2+ antagonists
.
J. Neurosci.
7
,
882
893
.
Niidome
,
T.
,
Kim
,
M. S.
,
Friedrich
,
T.
and
Mori
,
Y.
(
1992
).
Molecular cloning and characterization of a novel calcium channel from rabbit brain
.
FEBS Lett.
308
,
7
13
.
Okamura
,
Y.
(
1999
).
Functional expression of a protochordate L-type Ca2+ channel (abstract)
.
Biophys. J.
76
,
A340
.
Olcese
,
R.
,
Qin
,
N.
,
Schneider
,
T.
,
Neely
,
A.
,
Wei
,
X.
,
Stefani
,
E.
and
Birnbaumer
,
L.
(
1994
).
The amino terminus of a calcium channel beta subunit sets rates of channel inactivation independently of the subunit’s effect on activation
.
Neuron
13
,
1433
1438
.
Olivera
,
B. M.
,
Miljanich
,
G. P.
,
Ramachandran
,
J.
and
Adams
,
M. E.
(
1994
).
Calcium channel diversity and neurotransmitter release: the ω-conotoxins and ω-agatoxins
.
Annu. Rev. Biochem.
63
,
823
867
.
Ono
,
K.
,
Suga
,
H.
,
Iwabe
,
N.
,
Kuma
,
K.
and
Miyata
,
T.
(
1999
).
Multiple protein tyrosine phosphatases in sponges and explosive gene duplication in the early evolution of animals before the parazoan–eumetazoan split
.
J. Mol. Evol.
48
,
654
662
.
Paidhungat
,
M.
and
Garrett
,
S.
(
1997
).
A homolog of mammalian, voltage-gated calcium channels mediates yeast pheromonestimulated Ca2+ uptake and exacerbates the cdc1(Ts) growth defect
.
Mol. Cell Biol.
17
,
6339
6347
.
Papazian
,
D. M.
,
Schwarz
,
T. L.
,
Tempel
,
B. L.
,
Jan
,
Y. N.
and
Jan
,
L. Y.
(
1987
).
Cloning of genomic and complementary DNA from Shaker, a putative potassium channel gene from Drosophila
.
Science
237
,
749
753
.
Patton
,
D. E.
,
Silva
,
T.
and
Bezanilla
,
F.
(
1997
).
RNA editing generates a diverse array of transcripts encoding squid Kv2 K+ channels with altered functional properties
.
Neuron
19
,
711
722
.
Peixoto
,
A. A.
,
Smith
,
L. A.
and
Hall
,
J. C.
(
1997
).
Genomic organization and evolution of alternative exons in a Drosophila calcium channel gene
.
Genetics
145
,
1003
1013
.
Pelzer
,
S.
,
Barhanin
,
J.
,
Pauron
,
D.
,
Trautwein
,
W.
,
Lazdunski
,
M.
and
Pelzer
,
D.
(
1989
).
Diversity and novel pharmacological properties of Ca2+ channels in Drosophila brain membranes
.
EMBO J.
8
,
2365
2371
.
Perez-Reyes
,
E.
,
Castellano
,
A.
,
Kim
,
H. S.
,
Bertrand
,
P.
,
Baggstrom
,
E.
,
Lacerda
,
A. E.
,
Wei
,
X.
and
Birnbaumer
,
L.
(
1992
).
Cloning and expression of a cardiac/brain β subunit of the L-type calcium channel
.
J. Biol. Chem.
267
,
1792
1797
.
Perez-Reyes
,
E.
,
Cribbs
,
L. L.
,
Daud
,
A.
,
Lacerda
,
A. E.
,
Barclay
,
J.
,
Williamson
,
M. P.
,
Fox
,
M.
,
Rees
,
M.
and
Lee
,
J. H.
(
1998
).
Molecular characterization of a neuronal low-voltage-activated T-type calcium channel
.
Nature
391
,
896
900
.
Perez-Reyes
,
E.
and
Schneider
,
T.
(
1995
).
Molecular biology of calcium channels
.
Kidney Int.
48
,
1111
1124
.
Peterson
,
B. Z.
,
Tanada
,
T. N.
and
Catterall
,
W. A.
(
1996
).
Molecular determinants of high affinity dihydropyridine binding in L-type calcium channels
.
J. Biol. Chem.
271
,
5293
5296
.
Pichler
,
M.
,
Cassidy
,
T. N.
,
Reimer
,
D.
,
Haase
,
H.
,
Kraus
,
R.
,
Ostler
,
D.
and
Striessnig
,
J.
(
1997
).
β subunit heterogeneity in neuronal L-type Ca2+ channels
.
J. Biol. Chem.
272
,
13877
13882
.
Pragnell
,
M.
,
De Waard
,
M.
,
Mori
,
Y.
,
Tanabe
,
T.
,
Snutch
,
T. P.
and
Campbell
,
K. P.
(
1994
).
Calcium channel β-subunit binds to a conserved motif in the I–II cytoplasmic linker of the α1-subunit
.
Nature
368
,
67
70
.
Randall
,
A. D.
and
Tsien
,
R. W.
(
1997
).
Contrasting biophysical and pharmacological properties of T-type and R-type calcium channels
.
Neuropharmac.
36
,
879
893
.
Ranganathan
,
R.
(
1994
).
Evolutionary origins of ion channels
.
Proc. Natl. Acad. Sci. USA
91
,
3484
3486
.
Ren
,
D.
,
Xu
,
H.
,
Eberl
,
D. F.
,
Chopra
,
M.
and
Hall
,
L. M.
(
1998
).
A mutation affecting dihydropyridine-sensitive current levels and activation kinetics in Drosophila muscle and mammalian heart calcium channels
.
J. Neurosci.
18
,
2335
2341
.
Richmond
,
J. E.
,
Penner
,
R.
,
Keller
,
R.
and
Cooke
,
I. M.
(
1996
).
Characterization of the Ca2+ current in isolated terminals of crustacean peptidergic neurons
.
J. Exp. Biol.
199
,
2053
2059
.
Rosenthal
,
J. J.
and
Gilly
,
W. F.
(
1993
).
Amino acid sequence of a putative sodium channel expressed in the giant axon of the squid Loligo opalescens
.
Proc. Natl. Acad. Sci. USA
90
,
10026
10030
.
Ruth
,
P.
,
Röhrkasten
,
A.
,
Biel
,
M.
,
Bosse
,
E.
,
Regulla
,
S.
,
Meyer
,
H. E.
,
Flockerzi
,
V.
and
Hofmann
,
F.
(
1989
).
Primary structure of the β subunit of the DHP-sensitive calcium channel from skeletal muscle
.
Science
245
,
1115
1118
.
Saitou
,
N.
and
Nei
,
M.
(
1987
).
The neighbor-joining method: a new method for reconstructing phylogenetic trees
.
Mol. Biol. Evol.
4
,
406
425
.
Salkoff
,
L.
,
Butler
,
A.
,
Wei
,
A.
,
Scavarda
,
N.
,
Giffen
,
K.
,
Ifune
,
C.
,
Goodman
,
R.
and
Mandel
,
G.
(
1987
).
Genomic organization and deduced amino acid sequence of a putative sodium channel gene in Drosophila
.
Science
237
,
744
748
.
Salkoff
,
L.
and
Jegla
,
T.
(
1995
).
Surfing the DNA databases for K+ channels nets yet more diversity
.
Neuron
15
,
489
492
.
Sato
,
C.
and
Matsumoto
,
G.
(
1992
).
Primary structure of squid sodium channel deduced from the complementary DNA sequence
.
Biochem. Biophys. Res. Commun.
186
,
61
68
.
Schäfer
,
S.
,
Rosenboom
,
H.
and
Menzel
,
R.
(
1994
).
Ionic currents of Kenyon cells from the mushroom body of the honeybee
.
J. Neurosci.
14
,
4600
4612
.
Schafer
,
W. R.
and
Kenyon
,
C. J.
(
1995
).
A calcium-channel homologue required for adaptation to dopamine and serotonin in Caenorhabditis elegans
.
Nature
375
,
73
78
.
Schafer
,
W. R.
,
Sanchez
,
B. M.
and
Kenyon
,
C. J.
(
1996
).
Genes affecting sensitivity to serotonin in Caenorhabditis elegans
.
Genetics
143
,
1219
1230
.
Scott
,
V. E. S.
,
De Waard
,
M.
,
Liu
,
H.
,
Gurnett
,
C. A.
,
Venzke
,
D. P.
,
Lennon
,
V. A.
and
Campbell
,
K. P.
(
1996
).
β subunit heterogeneity in N-type Ca2+ channels
.
J. Biol. Chem.
271
,
3207
3212
.
Simpson
,
L.
and
Emeson
,
R. B.
(
1996
).
RNA editing
.
Annu. Rev. Neurosci
.
19
,
27
52
.
Singer
,
D.
,
Biel
,
M.
,
Lotan
,
I.
,
Flockerzi
,
V.
,
Hofmann
,
F.
and
Dascal
,
N.
(
1991
).
The roles of the subunits in the function of the calcium channel
.
Science
253
,
1553
1557
.
Sinnegger
,
M. J.
,
Wang
,
Z.
,
Grabner
,
M.
,
Hering
,
S.
,
Striessnig
,
J.
,
Glossmann
,
H.
and
Mitterdorfer
,
J.
(
1997
).
Nine L-type amino acid residues confer full 1,4-dihydropyridine sensitivity to the neuronal calcium channel α1A subunit: role of L-type Met1188
.
J. Biol. Chem.
272
,
27686
27693
.
Skeer
,
J. M.
,
Norman
,
R. I.
and
Sattelle
,
D. B.
(
1996
).
Invertebrate voltage-dependent calcium channel subtypes
.
Biol. Rev.
71
,
137
154
.
Smith
,
L. A.
,
Peixoto
,
A. A.
,
Kramer
,
E. M.
,
Villella
,
A.
and
Hall
,
J. M.
(
1998
).
Courtship and visual defects of cacophony mutants reveal functional complexity of a calcium-channel α1 subunit in Drosophila
.
Genetics
149
,
1407
1426
.
Smith
,
L. A.
,
Wang
,
X.
,
Peixoto
,
A. A.
,
Neumann
,
E. K.
,
Hall
,
L. M.
and
Hall
,
J. C.
(
1996
).
A Drosophila calcium channel α1 subunit gene maps to a genetic locus associated with behavioral and visual defects
.
J. Neurosci.
16
,
7868
7879
.
Sommer
,
B.
,
Kohler
,
M.
,
Sprengel
,
R.
and
Seeburg
,
P. H.
(
1991
).
RNA editing in brain controls a determinant of ion flow in glutamate-gated channels
.
Cell
67
,
11
19
.
Soong
,
T. W.
,
Stea
,
A.
,
Hodson
,
C. D.
,
Dubel
,
S. J.
,
Vincent
,
S. R.
and
Snutch
,
T. P.
(
1993
).
Structure and functional expression of a member of the low voltage-activated calcium channel family
.
Science
260
,
1133
1136
.
Spafford
,
J. D.
,
Spencer
,
A. N.
and
Gallin
,
W. J.
(
1998
).
A putative voltage-gated sodium channel α subunit (PpSCN1) from the hydrozoan jellyfish, Polyorchis penicillatus: structural comparisons and evolutionary considerations
.
Biochem. Biophys. Res. Commun.
244
,
772
780
.
Starr
,
T. V.
,
Prystay
,
W.
and
Snutch
,
T. P.
(
1991
).
Primary structure of a calcium channel that is highly expressed in the rat cerebellum
.
Proc. Natl. Acad. Sci. USA
88
,
5621
5625
.
Stea
,
A.
,
Soong
,
T. W.
and
Snutch
,
T. P.
(
1995
).
Voltage-gated calcium channels
. In
Handbook of Receptors and Channels: Ligand- and Voltage-Gated Ion Channels
(ed.
R. A.
North
), pp.
113
151
. Boca Raton, FL: CRC Press.
Strom
,
T. M.
,
Nyakatura
,
G.
,
Apfelstedt-Sylla
,
E.
,
Hellebrand
,
H.
,
Lorenz
,
B.
,
Weber
,
B. H.
,
Wutz
,
K.
,
Gutwillinger
,
N.
,
Ruther
,
K.
,
Drescher
,
B.
,
Sauer
,
C.
,
Zrenner
,
E.
,
Meitinger
,
T.
,
Rosenthal
,
A.
and
Meindl
,
A.
(
1998
).
An L-type calcium-channel gene mutated in incomplete X-linked congenital stationary night blindness
.
Nature Genetics
19
,
260
263
.
Suga
,
H.
,
Koyanagi
,
M.
,
Hoshiyama
,
D.
,
Ono
,
K.
,
Iwabe
,
N.
,
Kuma
,
K.
and
Miyata
,
T.
(
1999
).
Extensive gene duplication in the early evolution of animals before the parazoan–eumetazoan split demonstrated by G proteins and protein tyrosine kinases from sponge and hydra
.
J. Mol. Evol.
48
,
646
653
.
Sukharev
,
S. I.
,
Blount
,
P.
,
Martinac
,
B.
,
Blattner
,
F. R.
and
Kung
,
C.
(
1994
).
A large-conductance mechanosensitive channel in E. coli encoded by mscL alone
.
Nature
368
,
265
268
.
Tanabe
,
T.
,
Beam
,
K. G.
,
Adams
,
B. A.
,
Niidome
,
T.
and
Numa
,
S.
(
1990
).
Regions of the skeletal muscle dihydropyridine receptor critical for excitation–contraction coupling
.
Nature
346
,
567
569
.
Tanabe
,
T.
,
Takeshima
,
H.
,
Mikami
,
A.
,
Flockerzi
,
V.
,
Takahashi
,
H.
,
Kangawa
,
K.
,
Kojima
,
M.
,
Matsuo
,
H.
,
Hirose
,
T.
and
Numa
,
S.
(
1987
).
Primary structure of the receptor for calcium channel blockers from skeletal muscle
.
Nature
328
,
313
318
.
Tang
,
S.
,
Yatani
,
A.
,
Bahinski
,
A.
,
Mori
,
Y.
and
Schwartz
,
A.
(
1993
).
Molecular localization of regions in the L-type calcium channel critical for dihydropyridine action
.
Neuron
11
,
1013
1021
.
Tareilus
,
E.
,
Roux
,
M.
,
Qin
,
N.
,
Olcese
,
R.
,
Zhou
,
J.
,
Stefani
,
E.
and
Birnbaumer
,
L.
(
1997
).
A Xenopus oocyte β subunit: evidence for a role in the assembly/expression of voltage-gated calcium channels that is separate from its role as a regulatory subunit
.
Proc. Natl. Acad. Sci. USA
94
,
1703
1708
.
The C. elegans Sequencing Consortium
. (
1998
).
Genome sequence of the nematode C. elegans: a platform for investigating biology
.
Science
282
,
2012
2018
.
Thompson
,
J. D.
,
Higgins
,
D. G.
and
Gibson
,
T. J.
(
1994
).
CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, position-specific gap penalties and weight matrix choice
.
Nucleic Acids Res.
22
,
4673
4680
.
Thuleau
,
P.
,
Ward
,
J. M.
,
Ranjeva
,
R.
and
Schroeder
,
J. I.
(
1994
).
Voltage-dependent calcium-permeable channels in the plasma membrane of a higher plant cell
.
EMBO J.
13
,
2970
2975
.
Trudeau
,
L. E.
,
Baux
,
G.
,
Fossier
,
P.
and
Tauc
,
L.
(
1993
).
Transmitter release and calcium currents at an Aplysia buccal ganglion synapse. I. Characterization
.
Neurosci.
53
,
571
580
.
Varadi
,
G.
,
Lory
,
P.
,
Schultz
,
D.
,
Varadi
,
M.
and
Schwartz
,
A.
(
1991
).
Acceleration of activation and inactivation by the β subunit of the skeletal muscle calcium channel
.
Nature
352
,
159
162
.
Vogel
,
S. S.
,
Smith
,
R. M.
,
Baibakov
,
B.
,
Ikebuchi
,
Y.
and
Lambert
,
N. A.
(
1999
).
Calcium influx is required for endocytotic membrane retrieval
.
Proc. Natl. Acad. Sci. USA
96
,
5019
5024
.
Walker
,
D.
and
De Waard
,
M.
(
1998
).
Subunit interaction sites in voltage-dependent Ca2+ channels: role in channel function
.
Trends Neurosci.
21
,
148
154
.
Wei
,
A.
,
Jegla
,
T.
and
Salkoff
,
L.
(
1996
).
Eight potassium channel families revealed by the C. elegans genome project
.
Neuropharmac.
35
,
805
829
.
Wei
,
X.
,
Perez-Reyes
,
E.
,
Lacerda
,
A. E.
,
Schuster
,
G.
,
Brown
,
A. M.
and
Birnbaumer
,
L.
(
1991
).
Heterologous regulation of the cardiac Ca2+ channel α1 subunit by skeletal muscle β and γ subunits: implications for the structure of cardiac L-type Ca2+ channels
.
J. Biol. Chem.
266
,
21943
21947
.
Westenbroek
,
R. E.
,
Hell
,
J. W.
,
Warner
,
C.
,
Dubel
,
S. J.
,
Snutch
,
T. P.
and
Catterall
,
W. A.
(
1992
).
Biochemical properties and subcellular distribution of an N-type calcium channel α1 subunit
.
Neuron
9
,
1099
1115
.
Westenbroek
,
R. E.
,
Sakurai
,
T.
,
Elliott
,
E. M.
,
Hell
,
J. W.
,
Starr
,
T. V.
,
Snutch
,
T. P.
and
Catterall
,
W. A.
(
1995
).
Immunochemical identification and subcellular distribution of the α1A subunits of brain calcium channels
.
J. Neurosci.
15
,
6403
6418
.
White
,
B. H.
and
Kaczmarek
,
L. K.
(
1997
).
Identification of a vesicular pool of calcium channels in the bag cell neurons of Aplysia californica
.
J. Neurosci.
17
,
1582
1595
.
White
,
B. H.
,
Nick
,
T. A.
,
Carew
,
T. J.
and
Kaczmarek
,
L. K.
(
1998a
).
Protein kinase C regulates a vesicular class of calcium channels in the bag cell neurons of Aplysia
.
J. Neurophysiol.
80
,
2514
2520
.
White
,
G. B.
,
Pfahnl
,
A.
,
Haddock
,
S.
,
Lamers
,
S.
,
Greenberg
,
R. M.
and
Anderson
,
P. A.
(
1998b
).
Structure of a putative sodium channel from the sea anemone Aiptasia pallida
.
Invert. Neurosci.
3
,
317
326
.
Wicher
,
D.
and
Penzlin
,
H.
(
1997
).
Ca2+ currents in central insect neurons: electrophysiological and pharmacological properties
.
J. Neurophysiol.
77
,
186
199
.
Williams
,
M. E.
,
Brust
,
P. F.
,
Feldman
,
D. H.
,
Patthi
,
S.
,
Simerson
,
S.
,
Maroufi
,
A.
,
McCue
,
A. F.
,
Veliçelebi
,
G.
,
Ellis
,
S. B.
and
Harpold
,
M. M.
(
1992a
).
Structure and functional expression of an ω-conotoxin-sensitive human N-type calcium channel
.
Science
257
,
389
395
.
Williams
,
M. E.
,
Feldman
,
D. H.
,
McCue
,
A. F.
,
Brenner
,
R.
,
Veliçelebi
,
G.
,
Ellis
,
S. B.
and
Harpold
,
M. M.
(
1992b
).
Structure and functional expression of α1, α2 and β subunits of a novel human neuronal calcium channel subtype
.
Neuron
8
,
71
84
.
Wright
,
S. N.
,
Brodwick
,
M. S.
and
Bittner
,
G. D.
(
1996
).
Presynaptic calcium currents at voltage-clamped excitor and inhibitor nerve terminals of crayfish
.
J. Physiol., Lond.
496
,
347
361
.
Yamoah
,
E. N.
and
Crow
,
T.
(
1994
).
Two components of calcium currents in the soma of photoreceptors of Hermissenda
.
J. Neurophysiol.
72
,
1327
1336
.
Yang
,
J.
,
Ellinor
,
P. T.
,
Sather
,
W. A.
,
Zhang
,
J.-F.
and
Tsien
,
R. W.
(
1993
).
Molecular determinants of Ca selectivity and ion permeation in L-type Ca channels
.
Nature
366
,
158
161
.
Yokoyama
,
C. T.
,
Westenbroek
,
R. E.
,
Hell
,
J. W.
,
Soong
,
T. W.
,
Snutch
,
T. P.
and
Catterall
,
W. A.
(
1995
).
Biochemical properties and subcellular distribution of the neuronal class E calcium channel α1 subunit
.
J. Neurosci.
15
,
6419
6432
.
Zamponi
,
G. W.
,
Bourinet
,
E.
,
Nelson
,
D.
,
Nargeot
,
J.
and
Snutch
,
T. P.
(
1997
).
Crosstalk between G proteins and protein kinase C mediated by the calcium channel α1 subunit
.
Nature
385
,
442
446
.
Zhang
,
J. F.
,
Randall
,
A. D.
,
Ellinor
,
P. T.
,
Horne
,
W. A.
,
Sather
,
W. A.
,
Tanabe
,
T.
,
Schwarz
,
T. L.
and
Tsien
,
R. W.
(
1993
).
Distinctive pharmacology and kinetics of cloned neuronal Ca2+ channels and their possible counterparts in mammalian CNS neurons
.
Neuropharmac.
32
,
1075
1088
.
Zheng
,
W.
,
Feng
,
G.
,
Ren
,
D.
,
Eberl
,
D. F.
,
Hannan
,
F.
,
Dubald
,
M.
and
Hall
,
L. M.
(
1995
).
Cloning and characterization of a calcium channel α1 subunit from Drosophila melanogaster with similarity to the rat brain type D isoform
.
J. Neurosci.
15
,
1132
1143
.
Zoccola
,
D.
,
Tambutte
,
E.
,
Senegas-Balas
,
F.
,
Michiels
,
J. F.
,
Failla
,
J. P.
,
Jaubert
,
J.
and
Allemand
,
D.
(
1999
).
Cloning of a calcium channel α1 subunit from the reef-building coral, Stylophora pistillata
.
Gene
227
,
157
167
.
Zühlke
,
R. D.
,
Pitt
,
G. S.
,
Deisseroth
,
K.
,
Tsien
,
R. W.
and
Reuter
,
H.
(
1999
).
Calmodulin supports both inactivation and facilitation of L-type calcium channels
.
Nature
399
,
159
162
.