Voltage-gated Na+ channels play important functional roles in the generation of electrical excitability in most vertebrate and invertebrate species. These channels are members of a superfamily that includes voltage-gated K+, voltage-gated Ca2+ and cyclic-nucleotide-gated channels. There are nine genes encoding voltage-gated Na+ channels in mammals, with a tenth homologous gene that has not been shown to encode a functional channel. Other vertebrate and invertebrate species have a smaller number of Na+ channel genes. The mammalian genes can be classified into five branches in a phylogenetic tree, and they are localized on four chromosomes. Four of the branches representing the four chromosomal locations probably resulted from the chromosomal duplications that led to the four Hox gene clusters. These duplications occurred close to the emergence of the first vertebrates. The fifth branch probably evolved from a separate ancestral Na+ channel gene. There are two branches in the invertebrate tree, although members of only one of those branches have been demonstrated to encode functional voltage-gated Na+ channels. It is possible that the other branch may have diverged, so that its members do not represent true voltage-gated Na+ channels. Vertebrate and invertebrate Na+ channels appear to be derived from a single primordial channel that subsequently evolved independently in the two lineages.

Voltage-gated Na+ channels play a critical role in controlling the electrical excitability of animal cells, being primarily responsible for the rising phase of the action potential. The electrophysiological properties of all voltage-gated Na+ channels are not identical, however, and small differences in these properties can have significant effects on the electrical excitability of the cell (for a review, see Hille, 2001). Na+ channels with different functional or pharmacological properties have been observed in different tissues of the same species and in different species by electrophysiological recording (Mandel, 1992). In addition, a variety of different Na+ channel isoforms have been cloned, functionally expressed and characterized (Goldin, 1999).

The Na+ channel consists of a highly processed α subunit of approximately 260 kDa and contains four homologous domains termed I–IV. Within each domain, there are six transmembrane segments called S1–S6, and a hairpin-like P loop between S5 and S6 that comprises part of the channel pore (Fig. 1). The α subunit is associated with accessory subunits in the tissues of certain species, such as the β subunits in mammals (Catterall, 1993) and the tipE subunits in flies (Feng et al., 1995). The purpose of this review is to examine the phylogenetic relationships among the different voltage-gated Na+ channels that have been identified thus far. The structure and functional characteristics of the isoforms will not be discussed in detail as these topics have recently been reviewed (Catterall, 2000; Goldin, 2001).

For the purpose of this review, the voltage-gated Na+ channels can be considered in three distinct categories. Mammalian Na+ channels represent the largest group of genes that have been identified and studied (Table 1). These isoforms have all been classified as members of a single gene family with nine members. These will be referred to using the nomenclature that was proposed recently (Goldin et al., 2000), which consists of the prefix Nav to indicate the principal permeating ion and the principal physiological regulator, followed by a number that indicates the gene subfamily (currently Nav1 is the only subfamily). The number following the decimal point identifies the specific channel isoform (e.g. Nav1.1). There is a tenth isoform termed Nax that has approximately 50 % sequence identity to the other mammalian Na+ channels. Because none of the Nax channels has been functionally expressed, it is possible that this gene does not encode a voltage-gated Na+ channel, which is why it has not been assigned a number. Representatives of all nine isoforms and Nax have been identified and characterized from humans and rats, with representatives of some of the isoforms from other species. Only full-length sequences of the mammalian Na+ channels have been used in this analysis.

The other two groups of Na+ channel that will be discussed include the non-mammalian vertebrate Na+ channels and the invertebrate Na+ channels (Table 2). There is no systematic nomenclature for these channels, and they have been named in a variety of ways. A consistent nomenclature has been adopted for the purposes of this review. The name consists of the first letters of the genus and species, followed by Nav to indicate the principal permeating ion and the principal physiological regulator, and a number that was either part of the original name or is arbitrary.

There are many fewer representatives of non-mammalian Na+ channel sequences. Only five full-length sequences have been determined for non-mammalian vertebrate Na+ channels, and each of these is from a separate species. Lopreato et al. (2001) have determined partial sequences for six genes in one species, Sternopygus macrurus. This is the first attempt to identify all the Na+ channel genes in a single non-mammalian vertebrate species, so these data are particularly informative with respect to phylogeny. Therefore, these sequences have been included in the current analysis.

More invertebrate Na+ channel genes have been identified, with 13 full-length sequences representing 11 species. However, more than one full-length sequence has been determined for only two invertebrate species, Blattella germanica and Halocynthia roretzi. Two sequences have been determined from Drosophila melanogaster, with one full-length (DmNav1) and one that is almost complete (DmNav2). Blackshaw et al. (1999) have attempted to identify all the Na+ channel genes in an invertebrate species, Hirudo medicinalis. They determined partial sequences for four distinct genes, and these sequences have been included in the analysis.

Voltage-gated Na+ channels are members of a superfamily that includes voltage-gated K+ channels, voltage-gated Ca2+ channels and cyclic-nucleotide-gated channels (Jan and Jan, 1992; Hille, 2001). The Na+ and Ca2+ channel members all contain four homologous domains, whereas the K+ and cyclic-nucleotide-gated channels in the superfamily consist of tetramers of single-domain subunits. On the basis of structure alone, it might be expected that Na+ and Ca2+ channels would have evolved from the single-domain channels. This is in fact the most widely accepted scheme, for the reasons reviewed by Hille (1987, 1988). Bacteria, like eukaryotic cells, typically have high intracellular K+ and low intracellular Ca2+ concentrations, but many have no requirement for Na+ or Cl. Consistent with this lack of necessity for Na+ or Cl, bacteria do not generally express either Na+ or Cl channels, whereas they clearly express K+ channels (Milkman, 1994). Therefore, it is likely that an ancestral primordial prokaryotic channel was more similar to the voltage-gated K+ channels (Ranganathan, 1994). In fact, it has been suggested that channels with a cytoplasmic gate, such as the cyclic-nucleotide-gated channels and mechanosensitive channels, may represent the primordial channel from which the voltage-gated channels evolved (Jan and Jan, 1994; Anderson and Greenberg, 2001).

It is likely that the Ca2+ channels evolved from the K+ channels during the evolution of the stem eukaryotes (Hille, 1989). Protozoans generally use Ca2+ as the inward charge carrier, and purely Na+-dependent action potentials are not common until the cnidarians (Hille, 1989; Anderson and Greenberg, 2001). Na+ channels have not been detected in protozoa, algae or higher plants (Hille, 1989). According to this scenario, the Na+ channels evolved early in the metazoan era from an ancestral channel resembling the T-type (Cav3 family) Ca2+ channels, before the separation of diploblasts and triploblasts (Spafford et al., 1999).

The four domains of the Na+ channel have more similarity to the four domains of the Ca2+ channels than to each other, further supporting the argument that Na+ channels evolved after the subunit duplications leading to the Ca2+ channels (Hille, 1989; Strong et al., 1993). On the basis of sequence similarities, Strong et al. (1993) suggested that the original duplication event resulted in a two-domain channel consisting of domains I/III and II/IV, each of which then duplicated to result in the first four-domain Ca2+ channel. In this context, it is interesting to note that no channels consisting of two homologous domains have been observed. In addition, no K+ channels with multiple homologous domains have been observed, suggesting either that there was strong selective pressure against this form of channel or that the selectivity change from K+ to Ca2+ occurred before the gene duplication event (Anderson and Greenberg, 2001). The primordial Na+ channel then evolved from the Ca2+ channels, and this primordial Na+ channel subsequently evolved independently in vertebrates and invertebrates (Fig. 2) (Strong et al., 1993).

The mammalian vertebrate voltage-gated Na+ channels represent the best-characterized family in that all the different isoforms have been identified and cloned as full-length channels from two species, human and rat. The isoforms are in five distinct branches on the phylogenetic tree (Fig. 2). This division correlates well with the chromosomal localization of the genes encoding the channels, as described by Plummer et al. (1999). The genes encoding Nav1.1, Nav1.2, Nav1.3, Nav1.7 and Nax are clustered together on chromosome 2 in humans and mice (Table 1), and these channels form the largest branch of the tree. Nav1.1, Nav1.2, Nav1.3 and Nav1.7 share a number of common characteristics, including expression in the nervous system and block by nanomolar concentrations of tetrodotoxin (TTX). Nax is unusual in that this channel has sequence differences in functionally important regions, including fewer charges in the S4 voltage sensors and the absence of the IFM motif that is critical for fast inactivation (Fig. 1) (George et al., 1992a; Goldin, 2001). In addition, this is the only mammalian isoform that has not been functionally expressed in an exogenous system. Consistent with these differences, the gene for Nax has diverged more than any of the other channels in this branch of the tree. This divergence suggests that Nax may have evolved a unique functional role. In support of this hypothesis, mice in which the gene encoding Nax has been knocked out demonstrate functional deficits in salt intake behavior, suggesting that this channel has a physiological function other than as a voltage-gated Na+ channel (Watanabe et al., 2000).

The other mammalian Na+ channel genes that are clustered together are those encoding the Nav1.5, Nav1.8 and Nav1.9 isoforms. These genes are clustered on chromosome 9 in humans and the orthologous region of chromosome 3 in mice (Table 1). These isoforms also share functional characteristics, including the fact that they are all considered ‘TTX-resistant’ Na+ channels, requiring micromolar concentrations of TTX before they are blocked. Although the gene for Nav1.9 is clustered on the chromosome with the other two genes, this channel appears to represent a distinct branch of the tree, and it may have evolved from a distinct ancestral Na+ channel gene (Plummer and Meisler, 1999).

The final two isoforms, Nav1.4 and Nav1.6, each represent a separate branch of the tree, and the genes encoding these channels are located on separate chromosomes. The gene for Nav1.4 is located on chromosome 11 in humans and chromosome 17 in mice, and the gene for Nav1.6 is located on chromosome 15 in humans and chromosome 12 in mice (Table 1). The properties of these two channels are generally similar to those of the channels expressed in the nervous system, including block by nanomolar concentrations of TTX. Nav1.6 is the most closely related channel to the branch containing the nervous system Na+ channels, and this isoform is also expressed in the nervous system. Nav1.4 is the primary channel expressed in adult skeletal muscle, and is unique in this regard.

Plummer et al. (1999) observed that each of the chromosome segments containing the Na+ channel genes represents one of the Hox gene clusters. They suggested that the initial expansion of the Na+ channel genes was associated with the genomic duplications that occurred after the divergence of prevertebrates from invertebrate chordates and resulted in the four mammalian Hox gene clusters. Later tandem duplications of the Na+ channel genes on chromosomes 2 and 9 in humans (chromosomes 2 and 3 in mice) resulted in the two clusters of Na+ channel genes. Nav1.9 was more distantly related to Nav1.5 and Nav1.8 in the analysis of Plummer et al. (1999), and it appears on a separate branch in the current analysis (Fig. 2). Plummer et al. (1999) hypothesized that the ancestral chromosome might have had two Na+ channel genes, one that represented the ancestor of Nav1.9 and a second that represented the ancestor of the other mammalian Na+ channels. This model assumes that the ancestor of Nav1.9 was subsequently lost from three of the four chromosome segments.

The analysis of the non-mammalian vertebrate Na+ channels that have been characterized thus far is consistent with that of the mammalian channels and provides further information about the timing of the duplication events. The most informative data were obtained by Lopreato et al. (2001), who cloned partial sequences for six separate isoforms from the teleost fish Sternopygus macrurus. These sequences probably represent all the Na+ channels in that species. As pointed out by Lopreato et al. (2001), these isoforms align well with the branches of the mammalian channels (Fig. 2). SmNav3 and SmNav4 are in the branch with Nav1.1, SmNav5 aligns with Nav1.6, and SmNav1 and SmNav6 are in the branch with Nav1.4. SmNav2 forms a distinct branch of its own, although it is more closely related to Nav1.5 and Nav1.8 than to the other Na+ channel isoforms. SmNav2 was located on a branch with Nav1.5, Nav1.8 and Nav1.9 in the analysis of Lopreato et al. (2001). This branching pattern is not as robust as that for the other vertebrate channels though, because it is based on partial sequence data. Lopreato et al. (2001) suggested that the initial Na+ channel duplication into four genes occurred early in vertebrate history close to the emergence of the first vertebrates, so that a common ancestor of mammals and teleost fish already had four distinct genes. This hypothesis is consistent with the conclusions of Plummer et al. (1999). The duplications occurred as an initial event leading to two pairs of Hox genes, each of which then duplicated to result in the four clusters that are now present. These initial chromosomal duplications were then followed by tandem duplications, as suggested by Plummer et al. (1999), and these duplications were independent in teleosts and mammals. The relationships between the mammalian and teleost genes shown in Fig. 2 are supported by the corresponding tissue expression patterns of the mammalian and teleost channels encoded within each cluster (Lopreato et al., 2001).

An additional means of generating diversity in Na+ channels is through post-transcriptional processing, including alternative splicing and RNA editing. Alternative splicing has been demonstrated for five isoforms present in the mammalian nervous system, Nav1.1, Nav1.2, Nav1.3, Nav1.6 and Nav1.7 (Ahmed et al., 1990; Sarao et al., 1991; Schaller et al., 1992; Gustafson et al., 1993; Belcher et al., 1995; Plummer et al., 1997, 1998; Dietrich et al., 1998; Oh and Waxman, 1998). The proportion of differentially spliced transcripts depends on various factors, including age of development, the tissue of origin and the presence of modulatory agents such as dibutyryl cyclic AMP (Gustafson et al., 1993; Plummer et al., 1997; Dietrich et al., 1998; Oh and Waxman, 1998). However, electrophysiological differences resulting from alternative splicing have only been demonstrated for two alternatively spliced forms of Nav1.6 (Dietrich et al., 1998), so it is not clear how much functional diversity in mammalian Na+ channels results from alternative splicing. No alternative splicing has been demonstrated for non-mammalian vertebrate Na+ channels, and no RNA editing has been demonstrated for any vertebrate Na+ channels. This is in contrast to the situation with the invertebrate Na+ channels, which will be discussed below.

The data concerning the invertebrate Na+ channels are much less complete than those for mammalian Na+ channels in that more than one full-length sequence has been determined from only two species, Blattella germanica and Halocynthia roretzi. A full-length sequence and a collection of sequences comprising most of the coding region have been cloned from Drosophila melanogaster, and it is likely that these represent the only Na+ channel genes in that species (Littleton and Ganetzky, 2000). Blackshaw et al. (1999) have made the most extensive effort to identify all the Na+ channel genes in an invertebrate species, having isolated four partial clones from Hirudo medicinalis. The data for these four species suggest that there are two branches of the Na+ channel tree for invertebrates (Fig. 2), although HrNav1 is somewhat divergent from the first branch. The existence of two invertebrate Na+ channel branches was previously suggested by Spafford et al. (1999). The two invertebrate branches are independent of the branching that occurred during vertebrate evolution (Fig. 2). This result is consistent with the conclusion of Plummer et al. (1999) that the vertebrate gene duplications occurred after the divergence of prevertebrates from invertebrate chordates. The four genes from Hirudo medicinalis align with two in each of the branches, suggesting that two additional gene duplication events occurred later during the evolution of this species.

It is difficult to evaluate the functional significance of the two branches of invertebrate Na+ channels. Functional expression has been reported for only three channels, DmNav1 (Feng et al., 1995), MdNav1 (Smith et al., 1997) and BgNav1 (Tan et al., 2001; Liu et al., 2001b), and all these channels are in the same branch of the phylogenetic tree. However, K. Dong and colleagues have preliminary evidence for functional expression of the BgNav2 channel (K. Dong, personal communication). They suggest that the properties of this channel are so unique that it is questionable whether it should be considered a voltage-gated Na+ channel. If BgNav2 does not represent a true voltage-gated Na+ channel, it would suggest that there is only one branch of voltage-gated Na+ channels in invertebrates and that the second branch has diverged to the point that these channels carry out unique functions. If this is the case, then the functional diversity of Na+ channels in many, if not most, invertebrate species would have to result from post-transcriptional regulatory events, as suggested by Liu et al. (2001a), unless additional Na+ channel genes remain to be discovered in these species.

The two major post-transcriptional processes that would increase Na+ channel diversity are alternative splicing and RNA editing. Alternative splicing of Na+ channels has been shown to occur in two invertebrate species, Drosophila melanogaster (Loughney et al., 1989; Thackeray and Ganetzky, 1995) and Blattella germanica (Liu et al., 2001a). Spafford et al. (1999) pointed out that 85 % of the intron splice junctions in a jellyfish Na+ channel (PpNav1) are also found in mammalian Na+ channels with similar locations, suggesting that the positions of the introns within the coding regions have been retained from a common ancestor. However, no alternatively spliced variants were detected in PpNav1, suggesting that alternative splicing was not a primordial means of generating diversity in Na+ channels. RNA editing has been demonstrated for some invertebrate Na+ channels, including DmNav1 from Drosophila melanogaster (Hanrahan et al., 2000; Reenan, 2001). Hanrahan et al. (2000) observed that two of the three characterized RNA editing sites in Drosophila melanogaster are also conserved in Drosophila virilis, suggesting that this process has been maintained throughout the 61–65 million years of divergence between these two species.

Considering the fact that invertebrate species appear to have only 2–4 Na+ channel genes, in contrast to the nine mammalian genes, it is likely that invertebrates and vertebrates use fundamentally different means of generating diversity in Na+ channel function. The major means of generating diversity in invertebrate Na+ channels may involve alternative splicing and RNA editing, whereas vertebrate Na+ channel diversity may result primarily from the presence of multiple genes.

Voltage-gated Na+ channels are encoded by a family of genes that have been highly conserved throughout evolution, which undoubtedly reflects the critical functional role of these proteins in regulating electrical excitability. Because the majority of Na+ channels have been cloned from mammalian species, there is most information about the diversity of mammalian Na+ channels. The recent identification of additional Na+ channels in non-mammalian vertebrates suggests that many of the duplication events leading to the different isoforms occurred early in vertebrate evolution. There is much less definitive information about the invertebrate Na+ channels. Although a significant number of genes have been identified and sequenced, the limited amount of functional information makes it difficult to determine whether those genes encode true voltage-gated Na+ channels. Future efforts to express the invertebrate channels should help to more clearly define the roles of Na+ channel genes in these species.

Fig. 1.

A schematic diagram of the voltage-gated Na+ channel α subunit. The four homologous domains are labeled I–IV, and the six transmembrane segments in each domain are labeled S1–S6. The critical motif of the fast inactivation particle (IFM) is shown on the cytoplasmic linker connecting domains III and IV. The channel is aligned with the extracellular side of the membrane on the top. The lengths that are shown for the N and C termini and the interdomain cytoplasmic loops are consistent with the sequence of the Nav1.2 mammalian Na+ channel, but these linkers (except for that between domains III and IV) vary greatly in length and sequence among the different Na+ channel isoforms. P, P loop.

Fig. 1.

A schematic diagram of the voltage-gated Na+ channel α subunit. The four homologous domains are labeled I–IV, and the six transmembrane segments in each domain are labeled S1–S6. The critical motif of the fast inactivation particle (IFM) is shown on the cytoplasmic linker connecting domains III and IV. The channel is aligned with the extracellular side of the membrane on the top. The lengths that are shown for the N and C termini and the interdomain cytoplasmic loops are consistent with the sequence of the Nav1.2 mammalian Na+ channel, but these linkers (except for that between domains III and IV) vary greatly in length and sequence among the different Na+ channel isoforms. P, P loop.

Fig. 2.

This proposed phylogenetic tree for the voltage-gated Na+ channel α subunits was generated using sequences of the channels listed in Tables 1 and 2. The mammalian channels are in blue, the non-mammalian vertebrate channels are in green and the invertebrate channels are in red. This unrooted tree represents the optimal tree based on parsimony analysis of nucleotide sequences. To perform the analysis, the amino acid sequences for all the isoforms were aligned using Clustal W (Thompson et al., 1994). The amino acid sequences in the alignments were then replaced with the published nucleotide sequences, and the nucleotide sequence alignments were subjected to analysis using the program PAUP* (Swofford, 1998). Divergent portions, including most of the terminal regions and the cytoplasmic loops between domains I and II and domains II and III, were excluded from the PAUP* analysis. The numbers at the nodes indicate the bootstrap values for 100 replications. When a number is not indicated, the bootstrap value was less than 50. The scale bar represents 500 substitutions. Channels and species are identified in Tables 1 and 2.

Fig. 2.

This proposed phylogenetic tree for the voltage-gated Na+ channel α subunits was generated using sequences of the channels listed in Tables 1 and 2. The mammalian channels are in blue, the non-mammalian vertebrate channels are in green and the invertebrate channels are in red. This unrooted tree represents the optimal tree based on parsimony analysis of nucleotide sequences. To perform the analysis, the amino acid sequences for all the isoforms were aligned using Clustal W (Thompson et al., 1994). The amino acid sequences in the alignments were then replaced with the published nucleotide sequences, and the nucleotide sequence alignments were subjected to analysis using the program PAUP* (Swofford, 1998). Divergent portions, including most of the terminal regions and the cytoplasmic loops between domains I and II and domains II and III, were excluded from the PAUP* analysis. The numbers at the nodes indicate the bootstrap values for 100 replications. When a number is not indicated, the bootstrap value was less than 50. The scale bar represents 500 substitutions. Channels and species are identified in Tables 1 and 2.

Table 1.
graphic
graphic
Table 2.
graphic
graphic

I thank Drs Miriam Meisler and Ke Dong for critical reading of the manuscript, Dr George Gutman for help with the phylogenetic analysis, and Drs Ke Dong, Bob Maue, Greg Lopreato and Harold Zakon for supplying sequence data prior to publication. Work in the author’s laboratory is supported by grants from the NIH and AHA.

Ahmed, C. M. I., Auld, V. J., Lester, H. A., Dunn, R. and Davidson, N. (
1990
). Both sodium channel II and IIA α subunits are expressed in rat brain.
Nucleic Acids Res
.
18
,
5907
.
Ahmed, C. M. I., Ware, D. H., Lee, S. C., Patten, C. D., Ferrer-Montiel, A. V., Schinder, A. F., McPherson, J. D., Wagner-McPherson, C. B., Wasmuth, J. J., Evans, G. A. and Montal, M. (
1992
). Primary structure, chromosomal localization and functional expression of a voltage-gated sodium channel from human brain.
Proc. Natl. Acad. Sci. USA
89
,
8220
–8224.
Akopian, A. N., Sivilotti, L. and Wood, J. N. (
1996
). A tetrodotoxin-resistant voltage-gated sodium channel expressed by sensory neurons.
Nature
379
,
257
–262.
Akopian, A. N., Souslova, V., Sivilotti, L. and Wood, J. N. (
1997
). Structure and distribution of a broadly expressed atypical sodium channel.
FEBS Lett
.
400
,
183
–187.
Ambrose, C., Cheng, S., Fontaine, B., Nadeau, J. H., MacDonald, M. and Gusella, J. F. (
1992
). The alpha-subunit of the skeletal muscle sodium channel is encoded proximal to Tk-1 on mouse chromosome 11.
Mammal. Genome
3
,
151
–155.
Anderson, P. A. V. and Greenberg, R. M. (
2001
). Phylogeny of ion channels: clues to structure and function.
Comp. Biochem. Physiol
.
129B
,
17
–28.
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.
Auld, V. J., Goldin, A. L., Krafte, D. S., Marshall, J., Dunn, J. M., Catterall, W. A., Lester, H. A., Davidson, N. and Dunn, R. J. (
1988
). A rat brain Na+ channel α subunit with novel gating properties.
Neuron
1
,
449
–461.
Beckers, M.-C., Ernst, E., Belcher, S., Howe, J., Levenson, R. and Gros, P. (
1996
). A new sodium channel α-subunit gene (Scn9a) from Schwann cells maps to the Scn1a, Scn2a, Scn3a cluster of mouse chromosome 2.
Genomics
36
,
202
–205.
Belcher, S. M., Zerillo, C. A., Levenson, R., Ritchie, J. M. and Howe, J. R. (
1995
). Cloning of a sodium channel α subunit from rabbit Schwann cells.
Proc. Natl. Acad. Sci. USA
92
,
11034
–11038.
Blackshaw, S. E., Gross, R. H., Porter, D. M., Henderson, L. P. and Maue, R. A. (
1999
). Cloning of a family of putative sodium channel alpha subunit cDNAs from the leech.
J. Physiol., Lond
.
518
,
126P
.
Burgess, D. L., Kohrman, D. C., Galt, J., Plummer, N. W., Jones, J. M., Spear, B. and Meisler, M. H. (
1995
). Mutation of a new sodium channel gene, Scn8a, in the mouse mutant ‘motor endplate disease’.
Nature Genet
.
10
,
461
–465.
Catterall, W. A. (
1993
). Structure and function of voltage-gated ion channels.
Trends Neurosci
.
16
,
500
–506.
Catterall, W. A. (
2000
). From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels.
Neuron
26
,
13
–25.
Chen, J., Ikeda, S. R., Lang, W., Isales, C. M. and Wei, X. (
1997
). Molecular cloning of a putative tetrodotoxin-resistant sodium channel from dog nodose ganglion neurons.
Gene
202
,
7
–14.
Chen, Y. H., Dale, T. J., Romanos, M. A., Whitaker, W. R. J., Xie, X. M. and Clare, J. J. (
2000
). Cloning, distribution and functional analysis of the type III sodium channel from human brain.
Eur. J. Neurosci
.
12
,
4281
–4289.
Dib-Hajj, S. D., Tyrrell, L., Black, J. A. and Waxman, S. G. (
1998
). NaN, a novel voltage-gated Na channel, is expressed preferentially in peripheral sensory neurons and down-regulated after axotomy.
Proc. Natl. Acad. Sci. USA
95
,
8963
–8968.
Dib-Hajj, S. D., Tyrrell, L., Cummins, T. R., Black, J. A., Wood, P. M. and Waxman, S. G. (
1999
a). Two tetrodotoxin-resistant sodium channels in human dorsal root ganglion neurons.
FEBS Lett
.
462
,
117
–120.
Dib-Hajj, S. D., Tyrrell, L., Escayg, A., Wood, P. M., Meisler, M. H. and Waxman, S. G. (
1999
b). Coding sequence, genomic organization and conserved chromosomal localization of the mouse gene scn11a encoding the sodium channel NaN.
Genomics
59
,
309
–318.
Dietrich, P. S., McGivern, J. G., Delgado, S. G., Koch, B. D., Eglen, R. M., Hunter, J. C. and Sangameswaran, L. (
1998
). Functional analysis of a voltage-gated sodium channel and its splice variant from rat dorsal root ganglion.
J. Neurochem
.
70
,
2262
–2272.
Dong, K. (
1997
). A single amino acid change in the para sodium channel protein is associated with knockdown-resistance (kdr) to pyrethroid insecticides in german cockroach.
Insect Biochem. Mol. Biol
.
27
,
93
–100.
Dyer, J. R., Johnston, W. L., Castellucci, V. F. and Dunn, R. J. (
1997
). Cloning and tissue distribution of the Aplysia Na+ channel α-subunit cDNA.
DNA Cell Biol
.
16
,
347
–356.
Escayg, A., MacDonald, B. T., Meisler, M. H., Baulac, S., Huberfeld, G., An-Gourfinkel, I., Brice, A., LeGuern, E., Moulard, B., Chaigne, D., Buresi, C. and Malafosse, A. (
2000
). Mutations of SCN1A, encoding a neuronal sodium channel, in two families with GEFS+2.
Nature Genet
.
24
,
343
–345.
Felipe, A., Knittle, T. J., Doyle, K. L. and Tamkun, M. M. (
1994
). Primary structure and differential expression during development and pregnancy of a novel voltage-gated sodium channel in the mouse.
J. Biol. Chem
.
269
,
30125
–30131.
Feng, G., Deak, P., Chopra, M. and Hall, L. M. (
1995
). Cloning and functional analysis of TipE, a novel membrane protein that enhances Drosophila para sodium channel function.
Cell
82
,
1001
–1011.
Gautron, S., Dos Santos, G., Pinto-Henrique, D., Koulakoff, A., Gros, F. and Berwald-Netter, Y. (
1992
). The glial voltage-gated sodium channel: cell- and tissue-specific mRNA expression.
Proc. Natl. Acad. Sci. USA
89
,
7272
–7276.
Gellens, M. E., George, A. L., Jr, Chen, L., Chahine, M., Horn, R., Barchi, R. L. and Kallen, R. G. (
1992
). Primary structure and functional expression of the human cardiac tetrodotoxin-insensitive voltage-dependent sodium channel.
Proc. Natl. Acad. Sci. USA
89
,
554
–558.
George, A. L., Jr, Knittle, T. J. and Tamkun, M. M. (
1992
a). Molecular cloning of an atypical voltage-gated sodium channel expressed in human heart and uterus: evidence for a distinct gene family.
Proc. Natl. Acad. Sci. USA
89
,
4893
–4897.
George, A. L., Jr, Knops, J. F., Han, J., Finley, W. H., Knittle, T. J., Tamkun, M. M. and Brown, G. B. (
1994
). Assignment of a human voltage-dependent sodium channel α-subunit gene (SCN6A) to 2q21-q23.
Genomics
19
,
395
–397.
George, A. L., Jr, Komisarof, J., Kallen, R. G. and Barchi, R. L. (
1992
b). Primary structure of the adult human skeletal muscle voltage-dependent sodium channel.
Ann. Neurol
.
31
,
131
–137.
George, A. L., Jr, Ledbetter, D. H., Kallen, R. G. and Barchi, R. L. (
1991
). Assignment of a human skeletal muscle sodium channel alpha-subunit gene (SCN4A) to 17q23.-25.3.
Genomics
9
,
555
–556.
George, A. L., Jr, Varkony, T. A., Drabkin, H. A., Han, J., Knops, J. F., Finley, W. H., Brown, G. B., Ward, D. C. and Haas, M. (
1995
). Assignment of the human heart tetrodotoxin-resistant voltage-gated Na+ channel α-subunit gene (SCN5A) to band 3p21.
Cytogenet. Cell Genet
.
68
,
67
–70.
Goldin, A. L. (
1999
). Diversity of mammalian voltage-gated sodium channels. In Molecular and Functional Diversity of Ion Channels and Receptors, vol. 868 (ed. B. Rudy, and P. Seeburg), pp. 38–50. New York: New York Academy of Sciences.
Goldin, A. L. (
2001
). Resurgence of sodium channel research.
Annu. Rev. Physiol
.
63
,
871
–894.
Goldin, A. L., Barchi, R. L., Caldwell, J. H., Hofmann, F., Howe, J. R., Hunter, J. C., Kallen, R. G., Mandel, G., Meisler, M. H., Berwald-Netter, Y., Noda, M., Tamkun, M. M., Waxman, S. G., Wood, J. N. and Catterall, W. A. (
2000
). Nomenclature of voltage-gated sodium channels.
Neuron
28
,
365
–368.
Gustafson, T. A., Clevinger, E. C., O’Neill, T. J., Yarowsky, P. J. and Krueger, B. K. (
1993
). Mutually exclusive exon splicing of type III brain sodium channel α subunit RNA generates developmentally regulated isoforms in rat brain.
J. Biol. Chem
.
268
,
18648
–18653.
Hanrahan, C. J., Palladino, M. J., Ganetzky, B. and Reenan, R. A. (
2000
). RNA editing of the Drosophila para Na+ channel transcript: evolutionary conservation and developmental regulation.
Genetics
155
,
1149
–1160.
Hille, B. (
1987
). Evolutionary origins of voltage-gated channels and synaptic transmission. In Synaptic Function (ed. G. M. Edelman, W. E. Gall and W. M. Cowan), pp. 163–176. New York: John Wiley & Sons.
Hille, B. (
1988
). Evolutionary origin of electrical excitability. In Cellular Mechanisms of Conditioning and Behavioral Plasticity (ed. C. D. Woody, D. L. Alkon and J. L. McGaugh), pp. 511–518. New York: Plenum Press.
Hille, B. (
1989
). Ionic channels: evolutionary origins and modern roles.
Q. J. Exp. Physiol
.
74
,
785
–804.
Hille, B. (
2001
). Ion Channels of Excitable Membranes. Third edition. Sunderland, MA: Sinauer Associates, Inc.
Ingles, P. J., Adams, P. M., Knipple, D. C. and Soderlund, D. M. (
1996
). Characterization of voltage-sensitive sodium channel gene coding sequences from insecticide-susceptible and knockdown-resistant house fly strains.
Insect Biochem. Mol. Biol
.
26
,
319
–326.
Jan, L. Y. and Jan, Y. N. (
1992
). Tracing the roots of ion channels.
Cell
69
,
715
–718.
Jan, L. Y. and Jan, Y. N. (
1994
). Potassium channels and their evolving gates.
Nature
371
,
119
–122.
Jeong, S.-Y., Goto, J., Hashida, H., Suzui, T., Ogata, K., Masuda, N., Hirai, M., Isahara, K., Uchiyama, Y. and Kanazawa, I. (
2000
). Identification of a novel human voltage-gated sodium channel α subunit gene, SCN12A.
Biochem. Biophys. Res. Commun
.
267
,
262
–270.
Jeziorski, M. C., Greenberg, R. M. and Anderson, P. A. V. (
1997
). Cloning of a putative voltage-gated sodium channel from the turbellarian flatworm Bdelloura candida.
Parasitology
115
,
289
–296.
Joho, R. H., Moorman, J. R., VanDongen, A. M. J., Kirsch, G. E., Silberberg, H., Schuster, G. and Brown, A. M. (
1990
). Toxin and kinetic profile of rat brain type III sodium channel expressed in Xenopus oocytes.
Mol. Brain Res
.
7
,
105
–113.
Kallen, R. G., Sheng, Z.-H., Yang, J., Chen, L., Rogart, R. B. and Barchi, R. L. (
1990
). Primary structure and expression of a sodium channel characteristic of denervated and immature rat skeletal muscle.
Neuron
4
,
233
–242.
Kayano, T., Noda, M., Flockerzi, V., Takahashi, H. and Numa, S. (
1988
). Primary structure of rat brain sodium channel III deduced from the cDNA sequence.
FEBS Lett
.
228
,
187
–194.
Klugbauer, N., Lacinova, L., Flockerzi, V. and Hofmann, F. (
1995
). Structure and functional expression of a new member of the tetrodotoxin-sensitive voltage-activated sodium channel family from human neuroendocrine cells.
EMBO J
.
14
,
1084
–1090.
Kozak, C. A. and Sangameswaran, L. (
1996
). Genetic mapping of the peripheral sodium channel genes, Scn9a and Scn10a, in the mouse.
Mammal. Genome
7
,
787
–792.
Litt, M., Luty, J., Kwak, M., Allen, L., Magenis, R. E. and Mandel, G. (
1989
). Localization of a human brain sodium channel gene (SCN2A) to chromosome 2.
Genomics
5
,
204
–208.
Littleton, J. T. and Ganetzky, B. (
2000
). Ion channels and synaptic organization: analysis of the Drosophila genome.
Neuron
26
,
35
–43.
Liu, Z., Chung, I. and Dong, K. (
2001
a). Alternative splicing of the BSC1 gene generates tissue-specific isoforms in the German cockroach.
Insect Biochem. Mol. Biol
.
31
,
703
–713.
Liu, Z., Tan, J., Valles, S. M. and Dong, K. (
2001
b). Synergistic interaction between two cockroach sodium channel mutations and a tobacco budworm sodium channel mutation in reducing channel sensitivity to a pyrethroid insecticide.
Insect Biochem. Mol. Biol
. (in press).
Lopreato, G. F., Lu, Y., Southwell, A., Atkinson, N., Hillis, D., Wilcox, T. and Zakon, H. H. (
2001
). Evolution and divergence of sodium channel genes in vertebrates.
Proc. Natl. Acad. Sci. USA
98
,
7588
–7592.
Loughney, K., Kreber, R. and Ganetzky, B. (
1989
). Molecular analysis of the para locus, a sodium channel gene in Drosophila.
Cell
58
,
1143
–1154.
Lu, C.-M., Han, J., Rado, T. A. and Brown, G. B. (
1992
). Differential expression of two sodium channel subtypes in human brain.
FEBS Lett
.
303
,
53
–58.
Malo, D., Schurr, E., Dorfman, J., Canfield, V., Levenson, R. and Gros, P. (
1991
). Three brain sodium channel alpha-subunit genes are clustered on the proximal segment of mouse chromosome 2.
Genomics
10
,
666
–672.
Malo, M. S., Blanchard, B. J., Andresen, J. M., Srivastava, K., Chen, X.-N., Li, X., Jabs, E. W., Korenberg, J. R. and Ingram, V. M. (
1994
a). Localization of a putative human brain sodium channel gene (SCN1A) to chromosome band 2q24.
Cytogenet. Cell Genet
.
67
,
178
–186.
Malo, M. S., Srivastava, K., Andresen, J. M., Chen, X.-N., Korenberg, J. R. and Ingram, V. M. (
1994
b). Targeted gene walking by low stringency polymerase chain reaction: assignment of a putative human brain sodium channel gene (SCN3A) to chromosome 2q24-31.
Proc. Natl. Acad. Sci. USA
91
,
2975
–2979.
Mandel, G. (
1992
). Tissue-specific expression of the voltage-sensitive sodium channel.
J. Membr. Biol
.
125
,
193
–205.
Milkman, R. (
1994
). An Escherichia coli homologue of eukaryotic potassium channel proteins.
Proc. Natl. Acad. Sci. USA
91
,
3510
–3514.
Miyazaki, M., Ohyama, K., Dunlap, D. Y. and Matsumura, F. (
1996
). Cloning and sequencing of the para-type sodium channel gene from susceptible and kdr-resistant German cockroaches (Blattella germanica) and house fly (Musca domestica).
Mol. Gen. Genet
.
252
,
61
–68.
Nagahora, H., Okada, T., Yahagi, N., Chong, J. A., Mandel, G. and Okamura, Y. (
2000
). Diversity of voltage-gated sodium channels in the ascidian larval nervous system.
Biochem. Biophys. Res. Commun
.
275
,
558
–564.
Noda, M., Ikeda, T., Kayano, T., Suzuki, H., Takeshima, H., Kurasaki, M., Takahashi, H. and Numa, S. (
1986
). Existence of distinct sodium channel messenger RNAs in rat brain.
Nature
320
,
188
–192.
Noda, M., Shimizu, S., Tanabe, T., Takai, T., Kayano, T., Ikeda, T., Takahashi, H., Nakayama, H., Kanaoka, Y., Minamino, N., Kangawa, K., Matsuo, H., Raftery, M. A., Hirose, T., Inayama, S., Hayashida, H., Miyata, T. and Numa, S. (
1984
). Primary structure of Electrophorus electricus sodium channel deduced from cDNA sequence.
Nature
312
,
121
–127.
Ogata, K., Jeong, S.-Y., Murakami, H., Hashida, H., Suzuki, T., Masuda, N., Hirai, M., Isahara, K., Uchiyama, Y., Goto, J. and Kanazawa, I. (
2000
). Cloning and expression study of the mouse tetrodotoxin-resistant voltage-gated sodium channel α subunit NaT/Scn11a.
Biochem. Biophys. Res. Commun
.
267
,
271
–277.
Oh, Y. and Waxman, S. G. (
1998
). Novel splice variants of the voltage-sensitive sodium channel alpha subunit.
NeuroReport
9
,
1267
–1272.
Okamura, Y., Ono, F., Okagaki, R., Chong, J. A. and Mandel, G. (
1994
). Neural expression of a sodium channel gene requires cell-specific interactions.
Neuron
13
,
937
–948.
Plummer, N. W., Galt, J., Jones, J. M., Burgess, D. L., Sprunger, L. K., Kohrman, D. C. and Meisler, M. H. (
1998
). Exon organization, coding sequence, physical mapping and polymorphic intragenic markers for the human neuronal sodium channel gene SCN8A.
Genomics
54
,
287
–296.
Plummer, N. W., McBurney, M. W. and Meisler, M. H. (
1997
). Alternative splicing of the sodium channel SCN8A predicts a truncated two-domain protein in fetal brain and non-neuronal cells.
J. Biol. Chem
.
272
,
24008
–24015.
Plummer, N. W. and Meisler, M. H. (
1999
). Evolution and diversity of mammalian sodium channel genes.
Genomics
57
,
323
–331.
Potts, J. F., Regan, M. R., Rochelle, J. M., Seldin, M. F. and Agnew, W. S. (
1993
). A glial-specific voltage-sensitive Na channel gene maps close to clustered genes for neuronal isoforms on mouse chromosome 2.
Biochem. Biophys. Res. Commun
.
197
,
100
–104.
Rabert, D. K., Koch, B. D., Ilnikca, M., Obernolte, R. A., Naylor, S. L., Herman, R. C., Eglen, R. M., Hunter, J. C. and Sangameswaran, L. (
1998
). A tetrodotoxin-resistant voltage-gated sodium channel from human dorsal root ganglia, hPN3/SCN10A.
Pain
78
,
107
–114.
Ramaswami, M. and Tanouye, M. A. (
1989
). Two sodium-channel genes in Drosophila: implications for channel diversity.
Proc. Natl. Acad. Sci. USA
86
,
2079
–2082.
Ranganathan, R. (
1994
). Evolutionary origins of ion channels.
Proc. Natl. Acad. Sci. USA
91
,
3484
–3486.
Reenan, R. A. (
2001
). A→I pre-mRNA editing in animals.
Trends Genet
.
17
,
53
–56.
Rogart, R. B., Cribbs, L. L., Muglia, L. K., Kephart, D. D. and Kaiser, M. W. (
1989
). Molecular cloning of a putative tetrodotoxin-resistant rat heart Na+ channel isoform.
Proc. Natl. Acad. Sci. USA
86
,
8170
–8174.
Rosenthal, J. J. C. 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.
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.
Sangameswaran, L., Delgado, S. G., Fish, L. M., Koch, B. D., Jakeman, L. B., Stewart, G. R., Sze, P., Hunter, J. C., Eglen, R. M. and Herman, R. C. (
1996
). Structure and function of a novel voltage-gated, tetrodotoxin-resistant sodium channel specific to sensory neurons.
J. Biol. Chem
.
271
,
5953
–5956.
Sangameswaran, L., Fish, L. M., Koch, B. D., Rabert, D. K., Delgado, S. G., Ilnikca, M., Jakeman, L. B., Novakovic, S., Wong, K., Sze, P., Tzoumaka, E., Stewart, G. R., Herman, R. C., Chan, H., Eglen, R. M. and Hunter, J. C. (
1997
). A novel tetrodotoxin-sensitive, voltage-gated sodium channel expressed in rat and human dorsal root ganglia.
J. Biol. Chem
.
272
,
14805
–14809.
Sarao, R., Gupta, S. K., Auld, V. J. and Dunn, R. J. (
1991
). Developmentally regulated alternative RNA splicing of rat brain sodium channel mRNAs.
Nucleic Acids Res
.
19
,
5673
–5679.
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.
Schaller, K. L., Krzemien, D. M., McKenna, N. M. and Caldwell, J. H. (
1992
). Alternatively spliced sodium channel transcripts in brain and muscle.
J. Neurosci
.
12
,
1370
–1381.
Schaller, K. L., Krzemien, D. M., Yarowsky, P. J., Krueger, B. K. and Caldwell, J. H. (
1995
). A novel, abundant sodium channel expressed in neurons and glia.
J. Neurosci
.
15
,
3231
–3242.
Smith, M. R., Smith, R. D., Plummer, N. W., Meisler, M. H. and Goldin, A. L. (
1998
). Functional analysis of the mouse Scn8a sodium channel.
J. Neurosci
.
18
,
6093
–6102.
Smith, T. J., Lee, S. H., Ingles, P. J., Knipple, D. C. and Soderlund, D. M. (
1997
). The L1014F point mutation in the house fly Vssc1 sodium channel confers knockdown resistance to pyrethroids.
Insect Biochem. Mol. Biol
.
27
,
807
–812.
Souslova, V. A., Fox, M., Wood, J. N. and Akopian, A. N. (
1997
). Cloning and characterization of a mouse sensory neuron tetrodotoxin-resistant voltage-gated sodium channel gene, Scn10a.
Genomics
41
,
201
–209.
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.
Spafford, J. D., Spencer, A. N. and Gallin, W. J. (
1999
). Genomic organization of a voltage-gated Na+ channel in a hydrozoan jellyfish: insights into the evolution of voltage-gated Na+ channel genes.
Receptors Channels
6
,
493
–506.
Strong, M., Chandy, K. G. and Gutman, G. A. (
1993
). Molecular evolution of voltage-sensitive ion channel genes: on the origins of electrical excitability.
Mol. Biol. Evol
.
10
,
221
–242.
Swofford, D. L.. (
1998
). PAUP*. Phylogenetic analysis using parsimony (*and other methods). Sunderland, MA: Sinauer Associates.
Tan, J., Liu, Z., Tsai, T., Valles, S. M., Goldin, A. L. and Dong, K. (
2001
). Novel sodium channel gene mutations in Blattella germanica reduce the sensitivity of expressed channels to deltamethrin.
Insect Biochem. Mol. Biol
. (in press).
Tate, S., Benn, S., Hick, C., Trezise, D., John, V., Mannion, R. J., Costigan, M., Plumpton, C., Grose, D., Gladwell, Z., Kendall, G., Dale, K., Bountra, C. and Woolf, C. J. (
1998
). Two sodium channels contribute to the TTX-R sodium current in primary sensory neurons.
Nature Neurosci
.
1
,
653
–655.
Thackeray, J. R. and Ganetzky, B. (
1995
). Conserved alternative splicing patterns and splicing signals in the Drosophila sodium channel gene para.
Genetics
141
,
203
–214.
Thompson, J. D., Higgins, D. G. and Gibson, T. J. (
1994
). CLUSTAL W: improving the sensitivity of progressive multiple sequence alignment through sequence weighting, positions-specific gap penalties and weight matrix choice.
Nucleic Acids Res
.
22
,
4673
–4680.
Toledo-Aral, J. J., Moss, B. L., He, Z.-J., Koszowski, A. G., Whisenand, T., Levinson, S. R., Wolf, J. J., Silos-Santiago, I., Halegoua, S. and Mandel, G. (
1997
). Identification of PN1, a predominant voltage-dependent sodium channel expressed principally in peripheral neurons.
Proc. Natl. Acad. Sci. USA
94
,
1527
–1532.
Trimmer, J. S., Cooperman, S. S., Tomiko, S. A., Zhou, J., Crean, S. M., Boyle, M. B., Kallen, R. G., Sheng, Z., Barchi, R. L., Sigworth, F. J., Goodman, R. H., Agnew, W. S. and Mandel, G. (
1989
). Primary structure and functional expression of a mammalian skeletal muscle sodium channel.
Neuron
3
,
33
–49.
Tsai, C.-W., Tseng, J.-J., Lin, S.-C., Chang, C.-Y., Wu, J.-L., Horng, J.-F. and Tsay, J.-H. (
2001
). Primary structure and developmental expression of zebrafish sodium channel Nav1.6 during neurogenesis.
DNA Cell Biol
.
20
,
249
–255.
Wang, J., Rojas, C. V., Zhou, J., Schwartz, L. S., Nicholas, H. and Hoffman, E. P. (
1992
). Sequence and genomic structure of the human adult skeletal muscle sodium channel alpha subunit gene on 17q.
Biochem. Biophys. Res. Commun
.
182
,
794
–801.
Watanabe, E., Fujikawa, A., Matsunaga, H., Yasoshima, Y., Sako, N., Yamamoto, T., Saegusa, C. and Noda, M. (
2000
). Nav2/NaG channel is involved in control of salt-intake behavior in the CNS.
J. Neurosci
.
20
,
7743
–7751.
Williamson, M. S., Martinez-Torres, D., Hick, C. A. and Devonshire, A. L. (
1996
). Identification of mutations in the housefly para-type sodium channel gene associated with knockdown resistance (kdr) to pyrethroid insecticides.
Mol. Gen. Genet
.
252
,
51
–60.
Yotsu-Yamashita, M., Nishimori, K., Nitanai, Y., Isemura, M., Sugimoto, A. and Yasumoto, T. (
2000
). Binding properties of 3H-PbTx-3 and 3H-saxitoxin to brain membranes and to skeletal muscle membranes of puffer fish Fugu pardalis and the primary structure of a voltage-gated Na+ channel α-subunit (fMNa1) from skeletal muscle of F. pardalis.
Biochem. Biophys. Res. Commun
.
267
,
403
–412.