may: ' LIBRARY 10°? Michigan State University This is to certify that the thesis entitled FUNCTIONAL AND PHARMACOLOGICAL CHARACTERIZATION OF SIXTY-FOUR PARA SODIUM CHANNEL VARIANTS FROM DROSOPHILA MELANOGASTER IN XENOPUS OOCYTES presented by Rachel Lee O’Donnell Olson has been accepted towards fulfillment of the requirements for the MS. degree in EntomoLogy KIM-W Major Professor’s Signature 314 M. I 2— Z 00" Date MSU is an Affirmative Action/Equal Opportunity Institution ..-—.-.-.-.-¢--.--.-.p.-.—o--.--o— _.-.--.----o.----...-.--- _»__v cw -——— c- —-‘ ' PLACE IN RETURN BOX to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p2/CIRC/DaIeDue.indd-p.1 FUNCTIONAL AND PHARMACOLOGICAL CHARACTERIZATION OF SIXTY-FOUR PARA SODIUM CHANNEL VARIANTS FROM DROSOPHILA MELANOGASTER IN XENOPUS OOCYTES By Rachel Lee O’Donnell Olson A THESIS Submitted to Michigan State University in partial fulfillment of the requirements for the degree of MASTER OF SCIENCE Department of Entomology 2006 ABSTRACT FUNCTIONAL AND PHARMACOLOGICAL CHARACTERIZATION OF SIXTY- FOUR PARA SODIUM CHANNEL VARIANTS IN XENOPUS OOCYTES By Rachel Lee O’Donnell Olson Extensive alternative splicing and RNA editing of para, the voltage-gated sodium channel of Drosophila melanogaster, has been documented. However, the functional consequences of these post-transcriptional modifications have not yet been investigated. In this study we report the isolation, expression, and functional Characterization of sixty-four full-length para CDNA clones. The sixty- four clones were grouped into twenty-nine splice types based on alternative exon usage. Functional analysis of these variants revealed a broad spectrum of voltage-dependences of activation and to a lesser extent, voltage dependences of inactivation. Sequence comparison of three functionally distinct variants uncovered novel A to I editing sites. Furthermore, differential sensitivities to an a-scorpion toxin, thorlT, were revealed. The observed functional and pharmacological diversity of para variants supports the notion that alternative splicing and RNA editing are the two major mechanisms for generating sodium channel diversity in insects. ACKNOWLEDGMENTS I would like to express my sincere thanks to Dr. Ke Dong for providing me with this opportunity, her guidance, support, and encouragement throughout the course of my master's research. I would also like to thank Dr. Robert Hollingworth, Dr. Suzanne Thiem, and Dr. Edward Walker for their guidance and support. Appreciation is also extended to the members of Dr. Ke Dong’s laboratory. I would like to thank Dr. Weizhong Song and Dr. Yuhze Du for teaching me electrophysiological techniques, Mrs. Yoshiko Nomura for her assistance and instruction in molecular biology techniques, and Dr. Zhiqi Liu for cloning sixty-four para variants providing the foundation for my thesis work. A special thanks is also extended to Dr. Kristopher Silver for his critical review of my thesis. I would finally like to express my appreciation for the support and encouragement I received during the course of this research from my husband Rory Olson, and my parents Ruth and Jim O’Donnell. iii TABLE OF CONTENTS List of Tables and Figures for Chapter 1 .............................................. vi List of Tables and Figures for Chapter 3VII Chapter 1: Literature Review .................................................................... 1 |. Discovery of Sodium Current .................................................... 1 ll. Isolation of the Sodium Channel Protein ..................................... 1 III. Molecular Biology of the Sodium Channel ................................... 2 “LA Cloning of the First a—Subunit cDNAs ......................................... 2 "LB Functional Expression of the or-Subunit ....................................... 2 INC The Molecular Basis of Sodium Channel Function ........................ 3 |||.C1 Outer and Inner Pores, Selectivity Filter ...................................... 3 |||.CZ S4 Segments; Activation ......................................................... 4 |||.CS IFMT Motif; Inactivation ........................................................... 5 IV. Mammalian Voltage-Gated Sodium Channels .............................. 7 WA Tissue-Specific Expression ...................................................... 7 NB Developmental-Specific Expression ........................................... 8 WC Unique Electrophysiological Gating Properties of Mammalian VGSCs ................................................................ 8 V. Insect Voltage-Gated Sodium Channel ....................................... 9 VA Isolation of the Drosophila Voltage-Gated Sodium Channel Gene....10 V.B Conservation of Alternative Splicing of the Insect VGSC Genes ...... 13 V.B1 D. melanogaster ................................................................... 13 v.32 D. vin’lis ............................................................................... 14 was Musca domestica ................................................................... 15 V.B4 Blattella germanica .................................................................. 15 V.C Alternative Splicing Generates Distinct Gating Properties of Insect Voltage-Gated Sodium Channels ..................................... 16 VD RNA Editing Generates Diversity in IVGSCs ............................... 19 V.E Electrophysiological Analysis of RNA Edited Variants of BgNav ....... 20 VI. Pharmacology of Voltage-Gated Sodium Channels ....................... 21 Chapter 2: Specific Aims, Rationale, and Significance ........................... 25-26 Chapter 3: Functional and Pharmacological Characterization of Sixty-Four Para Sodium Channel Variants in Xenopus Oocytes ........................................................................ 27 Abstract ........................................................................................ 27 Introduction ................................................................................... 28 Materials and Methods ..................................................................... 29 Synthesis of First Strand cDNAs ............................................... 29 PCR and Cloning of para full-length cDNAs ................................. 29 iv Expression of para channels in Xenopus Iaevis oocytes ................ 30 Electrophysiological Recording and Analysis ............................... 31 thaIT sensitivity assay ........................................................... 32 Results ......................................................................................... 33 Molecular Analysis of 64 full-length CDNA clones ......................... 33 Functional Expression of 64 Variants in Xenopus oocytes .............. 33 Sequence Analysis of Three Functionally Distinct Variants ............. 34 Examination of the sensitivity of Para variants to thalT ................ 39 Examination of the senstitivities of Para variantsto thalT ...................... 40 Discussion .................................................................................... 43 Functional Versus Non-Functional Para Variants .......................... 43 Identification of 29 para Splice Types ......................................... 45 Frequency of alternative splicing of insect sodium channels. . .. ........45 Identification of Three Novel RNA Editing Sites ........................... 46 Identification of High-Voltage Activated and Low-Voltage Activated Para Variants ................................................. 48 Identification of an thaIT-resistant Para variant .......................... 50 Literature Cited .............................................................................. 51 LIST OF TABLES AND FIGURES FOR CHAPTER 1 Figure 1. Table 1. Table 2. Figure 2. Table 3. Table 4. Topological transmembrane organization of the voltage-gated sodium channel ......................................... 4 Tissue-specific and developmental-specific expression patterns of the nine mammalian voltage—gated sodium channels .................................................................... 9 Gating properties of nine mammalian SCs .......................... 10 Topology of the insect SC containing positions of axons. . 14 Alignment of the predicted amino acid sequences for optional exons a, b, e, f, h, i, and j and of mutually exclusive exons old and k/l ............................................. 16-17 Neurotoxins that act on sodium channels ........................... 22 vi LIST OF TABLES AND FIGURES FOR CHAPTER 3 Figure 1. Table 1. Figure 2. Table 2. Table 3. Figure 3. Figure 4. Figure 5. Table 4. Table 5. Identification of 29 Splice Types ....................................... 35 Activation and inactivation properties of 32 variants ............ 36 Amino acid changes in variants 13, 18, and 50 .................... 38 Amino acid changes observed in variants 13, 18, and 50 ....... 39 Effects of 1 nM thdIT on para variants ............................ 40 Peak sodium current traces of variant # 8 before and after application of thalT ..................................................... 41 Dose response curves of variants 8 and 21 ........................ 41 Amino acid changes in variants 8 and 21 compared to GenBank # M32078 ...................................................... 42 Sequence comparison of variants 8 and 2143 Exon frequencies of Para, VSSC1, BgNav ......................... 47 vii CHAPTER 1: Literature Review I. Discovery of Sodium Current Action potentials are electrical signals by which excitable cells, such as neurons, receive, conduct, and transmit information. In a series of elegant experiments, Alan Hodgkin and Andrew Huxley demonstrated that action potentials are the result of ions moving through the plasma membrane (Hodgkin and Huxley, 1952 a, b). Furthermore, they established sodium as one of the ions responsible for the production of action potentials and hypothesized that sodium selective pores activate (open) and inactivate (close) in response to changes in membrane potential. Because of their significant discovery the Nobel Prize was appropriately awarded to Alan Hodgkin and Andrew Huxley. ll. Isolation of the Sodium Channel Protein The proteins that comprise the voltage-gated sodium channel (SC) have been purified from a variety of tissues and species (review by Catterall, 2000). Using the high affinity binding of the SC protein to a radioactive-labeled tetrodotoxin, the protein was first purified from the electric organ of the eel, Electrophorus electricus (Agnew et al. 1980). This purified SC consists of a single polypeptide of an approximate weight of 270 kDa. Subsequent photo- affinity labeling experiments using a sodium specific neurotoxin, a-scorpion toxin, led to the discovery of two peptides (36 and 260 kDa) from synaptosomal membranes of the rat brain. The isolated peptides are denoted as a (primary) and B (auxiliary) subunits of the SC. The a polypeptide corresponds to the 270 kDa polypeptide isolated from eel (Beneski and Catterall, 1980). Ill. Molecular Biology of the Sodium Channel With the development of molecular biology techniques, our understanding of the molecular biology of voltage-gated sodium channels has greatly advanced. The major achievements and discoveries are summarized below. III. A Cloning of the First (Jr-Subunit cDNAs Sodium channel oz-subunit cDNAs were isolated from electric eel by screening expression libraries of electroplax eel mRNA using oligonucleotides encoding the electric eel SC and antibodies against the eel SC protein (Noda et al., 1984). Sequencing the entire polypeptide from the deduced cDNAs revealed the primary amino acid structure of the SC. The secondary structure of the a—subunit was determined to have four homologous domains (Dl — DIV), each consisting of six transmembrane segments (S1 - S6) (Noda et al., 1984; review by Catterall, 2000). III. B Functional Expression of the (Ir-Subunit Expression of the pore-fonning a-subunit alone generates sodium currents in Xenopus oocytes and mammalian cell lines (Noda et al., 1986). However, channel gating and expression are modulated by the inclusion of 3 subunits and result in emulating channels in vivo (reviewed by Catterall, 2000). A total of four 3 subunits have been identified in mammals (B1, B2, 83, & 34) (Yu et al., 2003). These subunits are hypothesized to function as cell adhesion molecules involved in cell-to-cell interactions. It is important to note that the expression of oz and 3 subunits varies between tissues (reviewed by Isom, 2001). For example, the skeletal muscle sodium channel is a complex of a and B1 subunits, whereas, the mammalian brain is comprised of a, B1, and 32 subunits. III. C The Molecular Basis of Sodium Channel Function Molecular mechanisms pertaining to the structure-function relationship of mammalian SCs have been and are currently under investigation. Below is our current understanding of the molecular basis of SC function. Ill. C1 Outer and Inner Pores, Selectivity Filter The outer pore (extracellular opening) and selectivity filter (amino acids that confer the channel to be sodium selective) of the channel is formed by the membrane re-entrant loops connecting segments 5 and 6 of domains l-lV (Catterall, 2000). The neurotoxins tetrodotoxin ('l'l'X) and saxitoxin (STX) block the SC by binding to the outer pore and thereby preventing sodium ion entry into the cell (Hille, 1984). A pair of amino acids located in each reentrant loop was associated with TTX and STX binding along with forming the selectivity filter (Hille, 1984; Terlau et al., 1991 ). The outer most set of amino acids, EEDD, form the outermost ring of the channel and serve as the primary binding site of these neurotoxins, which led to the identification of the outer pore (Fig. 1). The amino acid residues DEKA form the selectivity filter in the innermost part of the outer pore (Fig. 1; Terlau et al., 1991). Mutating DEKA to EEEE causes the SC to become permeable to calcium ions (Heinemann et al., 1992). The inner pore (intracellular portion) of the SC is proposed to be formed by the 86 segments (Fig. 1; Striessnig et al., 1990), which is based on evidence that local anesthetics (LAs) enter and bind to residues in the S6 segments and block the pore from the intracellular side (Hille, 1984). Inactivation Inner Particle Pore Figure 1 A-B Topological transmembrane organlzation of the voltage-gated sodium channel. A. The a—subunit of the SC consists of four homologous domains (DI- DIV) each containing six transmembrane segments (S1-$6). Voltage-sensors are formed by the S4 segments in each domain (grey). The outer pore and selectivity filter are formed by the linker connecting 85-86 of all four domains (striped). The sodium- selective-filter is formed by the amino acids DEKA. The inactivation particle containing the IFMT motif is located on the D3-D4 loop. The inner pore is formed by the arrangement of the segments 5 and 6. B. Schematic transmembrane arrangement of Dl-DIV of the a-subunit forming the channel pore. The linker connecting 85-86 of each domain form the outer pore and contain the amino acids DEKA (top view). The S6 segments form the inner pore and are depicted by the bottom portion of the $6 cylinders being tilted inwards. lll. C2 S4 Segments; Activation A depolarization of the membrane electric field leads to activation of the sodium channel (Hodgkin and Huxley 1952a). Each S4 segment contains four to eight positively charged arginine or lysine residues separated from each other by two neutral residues. These positively charged S4 segments serve as a voltage sensor and respond to potential changes across the plasma membrane (Fig. 1; Noda et al. 1984). Two comparable models have been proposed to explain the mechanism underlying activation; the sliding helix and the helical screw (Catterall, 1986; Guy and Seetharamulu, 1986). Both models assume that the positive residues of the S4 segments form ion pairs with negatively charged residues located in neighboring segments. The binding of ion pairs are released upon membrane depolarization, causing a conformational change of the channel resulting in opening of the pore (Catterall, 1986; Guy and Seetharamulu, 1986). Mutating the positive residues of the voltage sensor to neutral residues altered the voltage-dependency of gating, providing strong evidence for the two proposed activation models (Stuhmer et al., 1989; Kontis et al., 1997). The outward movement of the S4 segments in response to depolarization was detected in an elegant set of cysteine scanning mutagenesis experiments, further establishing the role of S4 as voltage sensors in sodium channels (Refs. in Catterall, 2000). III. C3 The IFMT Motif; Inactivation. Inactivation of the sodium channel occurs within a few milliseconds after activation, closing the channel pore (Ulbricht, 2005). Site-directed antibody and site-directed mutagenesis studies revealed that the short intracellular linker connecting DIII to DIV plays an integral role in fast inactivation (Fig. 1) (Refs. in Catterall, 2000). A contiguous hydrophobic triad of isoleucine, phenylalanine, and methionine (IFM) positioned in the middle of this linker are crucial for fast inactivation (West et al., 1992). Site-directed anti-peptide antibodies to this linker (Vassilev et al., 1988, 1989), deletions of amino acids from this loop (Patton et al., 1992), as well as mutating the IFM triad to glutamate (QQQ) inhibited inactivation (West at al., 1992). Additional experiments demonstrated that peptides containing the IFM motif serve as pore blockers to inactivation deficient channels and completely restore inactivation, providing further evidence supporting the significant role of the IFM motif in inactivation (Eaholtz et al., 1994; McPhee 1998; Eaholtz et al., 1999). Recent experiments indicate that there is a fourth amino acid, threonine, in addition to the IFM triad, that is essential for fast inactivation (IFMT) (Rohl et al., 1999). Inactivation occurs with the occlusion of the pore resulting in the block of further ion entry into the cell (review by Goldin, 2003). The small linker connecting Dlll to DIV (containing the IFMT motif) is responsible for occlusion of the pore by binding to hydrophobic residues near the intracellular portion of the channel. Scanning mutagenesis experiments indicate that the docking site of the IFMT motif consists of the 84-85 linkers of DI" and DIV along with the cytoplasmic portion of DIVSB. It is hypothesized that the conformational change ensued by the S4 segments result in the exposure of the hydrophobic residues contained in the 84-85 segments which then interact with the IFMT motif, looking it into place. Binding of the IFMT motif to its docking site is thought to be facilitated by glycine and proline residues that flank the short intracellular linker connecting Dlll to DIV. Glycine residues confer flexibility in protein structure as proline residues offer structural rigidity and a bent conformation to the linker (refs. in Zhao et al., 2004). The flexibility of glycine and the structural rigidity of proline serve as‘ ‘molecular hinges’ for the small linker containing the IFMT motif. Thus, the linker connecting DIII to DIV has been described as a ‘hinged lid’ with the IFMT motif acting as a ‘latch’, locking the lid into place by binding to the intracellular portion of the pore and preventing further ion entry into the cell (described as a ‘hinged lid’ model; review by Catterall, 2000). IV. Mammalian Voltage-Gated Sodium Channels Voltage-gated sodium channel genes have been identified in various vertebrates and invertebrates such as humans, rats, leaches, fish, and several species of insects (review by Goldin, 2002). Nine different SC a-subunit genes, Nav1.1 — Nav1.9, have been identified in mammalian systems (Goldin et al. 2000; Catterall et al. 2003). These nine mammalian SC proteins share greater than 50% amino acid homology (Goldin et al. 2000). These 803 are expressed in different tissues and cell types and are localized in different sub-cellular compartments. Each has unique gating properties thought to fulfill distinctive roles in a given cell type. Mutations in the SC genes have been associated with various mammalian diseases including myoclonic epilepsy of infancy and hyperkalemic periodic paralysis (Claes et al., 2001; Kelly et al., 1997). NA Tissue-Specific Expression The nine isoforms of the mammalian SC exhibit distinct tissue distribution patterns. Transcripts of most isoforms are present in many different tissues, however," the abundance of each isoform is profound in only select tissues (review by Goldin, 1999). lsofonn Nav1.1 is abundantly expressed in the central nervous system (CNS) along with Nav1.2 and Nav 1.3. Additionally, Nav1.1 and Navt .3 are localized to the somas of neurons, while Navt .2 was detected in the nodes of Ranvier and in unmyelinated axons. Nav1.4 is specifically expressed in skeletal muscle and Nav1.5 is independently expressed in cardiac myocytes. lsoforrn Nav1.6 is most abundant in the CNS in comparison to the peripheral nervous system (PNS) but is also detected in dorsal root ganglia (DRG) neurons. Nav1.7 is mainly expressed in the PNS, being present in DRG neurons, Schwann cells, and in neuron-endocrine cells. The last two isoforms, Nav1.8 and Nav1.9, have been identified in the PNS DRG neurons and trigeminal ganglia. IV.B Developmental-Specific Expression Besides tissue-specific expression patterns, developmental specific- regulation of mammalian 805 have been documented for seven of the nine isoforms (Nav1.1 - Nav1.7; review by Goldin, 1999). Nav1.1 becomes detectible shortly after birth and increases in number until adulthood. Nav 1.2 is accumulated in embryonic nodes of Ranvier in some CNS tracts. The transcription level of isoform Nav1.3 peaks at birth and becomes undetectable in adulthood. Transcripts of Nav 1.5 are detectable in neonatal skeletal muscle but are replaced by Nav 1.4 in adult skeletal muscle. Nav 1.6 is the most abundantly expressed isoform found in the CNS during adulthood, but, is expressed in lower numbers during development. Transcripts of Nav 1.7 are located in the neonatal cortex of the CNS The remaining two isoforms, Nav 1.8 and Nav 1.9, have not yet been described to display developmental-specific expression patterns (review by Goldin, 1999). MC Unique Electrophysiological Gating Properties of Mammalian $08 The nine SC variants in mammals exhibit significant differences in functional properties. These differences likely contribute unique functional roles in distinct tissues or cell types. Specifically, the voltage required for activation and inactivation varies among the isoforms (Table 2). Table 1. Tissue-specific and developmental-specific expression patterns of the nine mammalian voltage-gated sodium channels SC Primary Tissue Developemental Type Distribution Profile Nav1.1 CNS adult Nav 1.2 CNS adult Nav 1.3 CNS fetus Nav 1.4 skeletal muscle adult Nav 1.5 heart muscle adult Nav 1.6 CNS adult Nav 1.7 PNS neonatal Nav 1.8 PNS Nav 1.9 PNS Table modified from Goldin, 1999 V. Insect Voltage-Gated Sodium Channel It is important to study insect 80$ for three key reasons. First, 803 are targeted sites of several Classes of insecticides, including DDT, pyrethroids, and indoxacarb (Soderlund, 2005). These insecticides are widely used in the control of agricultural and medically related pest species. However, resistance to current insecticides is a global problem. Mutations in the SC gene have already been shown to confer resistance to pyrethroids and DDT (Soderlund, 2005). Second, as presented below, the insect SC gene undergoes extensive post-transcriptional modifications including alternative splicing and RNA editing events. At the present time, we do not understand the contributions of alternative splicing and RNA editing in the regulation of neuronal excitability. Third, though there is great similarity in the structure-function relationships between the insect and mammalian channels, pharmacological differences have been reported (Cestele and Catterall, 2000). Thus, by better understanding the molecular basis of differential sensitivities of insect and mammalian $03 to various neurotoxins, valuable information will be provided toward the development of safe and selective insecticides. Table 2. Gating properties of nine mammalian SCs lsoforrn Activation (mV) Inactivation (mV) Notes Nav 1.1 -33 -72 Nav 1.2 -24 -53 Nav 1.3 -23 to -26 -65 to -69 Nav 1.4 -26 -50.1 or -56 -30 rat Nav 1.5 -47 or -56 -84 or -100 F as the major anion in intracellular solution -27 -61 aspartate as the major anion in intracellular solution Nav 1.6 -8.8 -55 cut-open oocyte clamp; mouse -17 -51 “ “ + B1& B2 -26 “ “ + inactivation removed -37.7 -97.6 macropatch clamp; rat Nav 1.7 -31 macropatch; rat -45 -65 TTX sensitive DRG neurons -78 two-electrode clamp; rat -60.5 whole-cell clamp -39.6V whole-cell clamp (+81) Nav 1.8 -16 to -21 ~-3O DRG neurons; rat Nav 1.9 -47 to -54 -44 to -54 DRG neurons; rat Modified from Catterall, Goldin, and Waxman (2005). VA Isolation of the Drosophila Voltage-Gated Sodium Channel Gene In 1987 the Qrosophila sodium channel 1 (DSC1) was isolated by screening a Drosophila genomic library with a CDNA probe encoding the eel SC. 10 The deduced amino acid sequence of DSC1 revealed high homology to that of the previously characterized vertebrate SCs. The secondary structure of DSC1 was deduced to contain four homologous domains, each containing six transmembrane segments like that of mammalian SCs. Importantly, the S4 segments of each domain of the 0801 channel also contained positive residues separated by two uncharged resides resembling the ‘voltage sensors’ described in mammalian sodium channels (Salkoff et al., 1987). In the same year, 1987, Ganetzky and colleagues isolated a second SC gene, para, by analyzing a temperature-sensitive paralytic phenotype of Drosophila mutants. These paralytic temperature-sensitive mutant flies undergo instantaneous paralysis when temperatures are elevated to 29°C but fully recover when restored to ambient temperature (22°C) (Suzuki et al. 1971; Siddiqi and Benzer, 1976). The observed paralysis was hypothesized to be the result of a temperature sensitive blockade of action potential conduction (Siddiqi and Benzer, 1976; Wu and Ganetzky, 1980). This para gene was isolated using P- element transposon tagging (Loughney et al., 1989). The complete Para protein was deduced from six overlapping CDNA clones and revealed four homologous domains like that described of other 80s as well as being highly homologous in the transmembrane regions (60%) to those of the rat brain channel (Loughney et al., 1989). In an independent study, the D. melanogaster genomic library was searched for sequences similar to rat brain 803. This study resulted in the isolation of partial cDNAs corresponding to para and DSC1, suggesting that there are two genes encoding SCs in insects (Ramaswami and Tanouye, 1989). 11 In situ hybridization and immunochemistry studies revealed distinct tissue and developmental distribution patterns of para and DSC1. Using para and DSC1 CDNA restriction fragments as probes, the expression patterns of para and DSC1 transcripts were analyzed in in-situ hybridization experiments. The results of this study indicate that para is ubiquitously expressed throughout the nervous system in all developmental stages, while DSC1 is detected in all stages except that of the embryonic stage (Hong and Ganetzky, 1994). Additionally, para is expressed throughout the CNS and PNS while DSC1 is restricted to a few cells in either system (Hong and Ganetzky, 1994; Castella et al., 2001 ). When functionally expressed the para transcripts were found to carry the inward voltage-dependent sodium current (Genneraad et al., 1992). Functional analysis of Para in Xenopus oocytes indicates that the biophysical properties of the Para channel are similar to those of mammalian SCs (Warmke et al., 1997). Also, various toxins that act on mammalian SCs also act on Para, examples include permethrin and deltamethrin (Vais et al., 2001). An ortholog of DSC1, BSC1, was cloned from the German cockroach, B. gennanica. The sequence homology between BSC1 and DSC1 is high, 81%, 78%, 80%, and 88% in domains I, II, III, and IV, respectively (Liu et al., 2001). The functional expression of BSC1 in Xenopus oocytes revealed that the BSC1 channel is more permeable to Ca2+ than to Na+ and exhibits unique gating properties different from those of sodium channels, indicating that BSC1 encodes a Cay-selective cation channel (Zhou et al., 2003, Liu et al., 2001). The functional analysis of DSC1 has not yet been published, however, personal communication with Dr. W. Song indicates that the DSC1 Channel is also 12 calcium-selective and possesses gating properties similar to those of the BSC1 channel (Dr. W. Song, a postdoctoral associate in Dr. K. Dong‘s laboratory). In summary, DSC1 does not encode a sodium channel, and para is the only SC gene in insects. V.B Conservation of Alternative Splicing of the Insect Sodium Channel Genes Because of the involvement of SC mutations in pyrethroid resistance in many important insect and arachnid pest species, CDNAs (partial or full-length) of the SC gene have been isolated from many arthropod species (Soderlund, 2005). However, alternative splicing and RNA editing of the insect SC has mainly been studied in four insect species: D. melanogaster, D. virillis, Musca domestica, and Blattella gennanica. V.B1 D. melanogaster: As mentioned previously, there exists a single SC gene, para, in D. melanogaster (Loughney et al., 1989; O’Dowd et al., 1989). Alternative splicing of para has been proposed to be the key mechanism influencing SC diversity (Loughney et al., 1989; O’Dowd and Aldrich, 1988; O’Dowd et al., 1995; Thackeray and Ganetzky, 1994, 1995; Warrnke et al., 1997). Seven alternative exons (a, b, e, f, h, i, j) and four mutually exclusive exons (old and k/l) have been identified (Loughney et al., 1989; O’Dowd et al., 1995; Thackeray and Ganetzky, 1994, 1995; Warrnke et al., 1997; Lee et al., 2002; Tan et. al., 2002a). The secondary structure of the insect SC along with the positioning of alternative splicing is presented in Figure 2. l3 In D. melanogaster alternative exons 0, e, and f have been shown to be developmentally regulated. The percentage of transcripts in embryos containing exons 0, e, and fwere 1.7%, 3.4% and 87%, respectively, but changed to 24%, 28% and 6.7%, respectively in adults (Thackeray and Ganetzky, 1994). It 3 l j i a Figure 2. Topology of the Insect SC containing positions of exons. Positions of sequences encoded by optional exons a, b, e, f, and h as well as mutually exclusive exons c and d in the sodium channel protein are highlighted in colored boxes. The letters below the boxes are the exons to which the boxes belong. d/c f h l/k V. 32 D. virilis The alternative exons of para identified in D. melanogaster are also found in D. virilis (Thackeray and Ganetzky, 1995). These alternative exons are highly conserved in amino acid sequence and are presented in Table 2 below. The protein sequence identities of the entire coding regions of D. virilis and D. melanogaster are 98.8% homologous, while the alternative exons are 99.5% homologous in amino acid structure (Thackeray and Ganetzky, 1995). Exons c, e, and f of D. virilis are developmentally regulated in a similar fashion to D. melanogaster. D. vin'Iis embryos contain 0%, 9.2% and 70.2% exons 0, e, and f, while adults contain 17.7%, 25%, and 11.3% respectively (Thackeray and Ganetzky, 1995). It was determined that 75% of D. melanogasater embryos and I4 57.3% of D. virilis embryos share the same exon usage; a+, b+, d, e-, f+ or a-, b+, d, e-, f+ (Thackeray and Ganetzky, 1994, 1995). V.B3 Musca domestica Like that of D. melanogaster and D. virilis, the Musca domestica (house fly) SC, Vssc1 (Voltage-sensitive sodium channel), contain exons a through h. In addition to these exons, two novel mutually exclusive exons k and l were identified in Vssc1. By searching the reported para sequence and performing a BLAST search exons k and I were subsequently identified in para (Lee et al., 2002). Exons b, f, h, and j of VSSC1 are identical in amino acid structure to Para, while exons a, d, e, and i have less than five changes and exon c has eight changes (Lee et al., 2002). Interestingly the regulation of alternative exons between D. melanogaster and M. domestica species is similar. Exons a, b, c, d, e, f, and i show comparable exon frequencies among the transcripts (Lee et al., 2002). The frequencies of exons h, j, k, and I among para transcripts have not yet analyzed. V.B4 Blattella germanica The cockroach SC, BgNav, is 78% and 76% identical in amino acid framework to Para and Vssc1 respectively (Dong, 1997). When comparing alternative exon sequences between D. melanogaster, D. virilis, M. domestica, and B. gennanica, the sequences from B. germanica are quite divergent (refer to Table 3). Exon a of BgNav contains fewer amino acids in comparison to exon a of para,, exon d is missing, and exons e, f, and i of BgNav variants contain more amino acid changes than D. virilis, and M. domestica when compared to para. It 15 is interesting to note that the three fly sequences are more closely related to each other than with the cockroach sequence. Three mutually exclusive exons have been identified in BgNav; G1, GZ, and G3. G1 is equivalent to exon l of para, G2 corresponds to exon k of para and GS is unique to BgNav, having not been identified in Para or other insect SCs. Exons Gt, GZ, and G3 of the cockroach are abundantly expressed in embryonic stages II and Ill, but were found in much lower frequencies in adults. Exon G1 is localized to the nerve cord, coxa and the remaining leg (femur, tibia, and tarsus). Exon G2 is the same as G1 but is expressed in the ovary and gut. Exon G3 is localized to the coxa and the remaining leg (Tan et al., 2002a). V.C Alternative Splicing Generates Distinct Gating Properties of Insect Voltage-Gated Sodium Channels Studies on mammalian 803 show that phosphorylation of protein kinase A (PKA) sites in the cytoplasmic linker connecting DI and DH modulates channel function (references in O’Dowd et al., 1995). The functional role of alternative splicing in para was initially investigated by O’Dowd and colleagues (O’Dowd et al., 1995). Table 3. Alignment of the predlcted amino acid sequences for optlonal exons a, b, e, f, h, i, and j and mutually exclusive exons cld and kll. Asterisks represent sequence homology and gaps were introduced for optimal alignment (dashes). The para and Vssc1 amino acid sequences were obtained from Lee et al., 2002, and correspond to GenBank accession #s M32078 and X96668, respectively. The D. virilis sequences (Thackeray and Ganetzky, 1994) corresponds to GenBank accession #s U26343 and U26722. The cockroach sequences were obtained from Song et al., 2004 and Tan et al., 2002, and correspond to the GenBank sequence U73583. Exon d has not been identified in BgNav (Personal communication with Mrs. Nomura in Ke Dong's laboratory). 16 Table. 3 Amino acid sequence alingent of optional exons he 0 Pan! VCscf Duflfifis BgNav Pan! Vksc1 DuWfifis BgNav Pan: VCsc1 Dufififls BgNav Pant thwc1 Duflfifis BgNav Pan! VCscf Duflflfis BgNav Pan! vssc1 Dufififis BgNav Fan! V%301 Duflflfis BgNav Pant Vksc1 Dufififis BgNav Pan! Vksc1 Duflfifis BgNav Pan: V%sc1 Duflfifls BgNav Pan: VCscf Duflfifls BgNav TSLSLPGSPPNLRRGSRSSHK ***********I********* ********************* A*****(a) [truncated exon] VSVYYPPT ******** ******** **I***** LRVFKLAKSWPTLNLLISIMGRTVGALGNLTFVLCIIIFIFAVMGMOLFGXNYTD *************************************************VFY8VT ******************************************************* *****************************************************y* LRVFKLAKSWPTLNLLISIMGRTMGALGNLTPVLCIIIPIFAVMGMQLFGKNYHD ****it***********************************************I* *‘k***‘k*'k*********************************************** GBRTNQISWINSB ***I********* ***I********* **_-**3*-swx* -GKGVCRCISA -********** -********** DARE—RDTDLDLT HIGHSINHQDNRLBHILNGRGLSIQ ******ag**********n****** ******************n****** VSVI-Q-RQPAPTTAHQAT-KVR-KV8T ****-*_********p-**_***-**** ****-*-********p-**-***-**** VP*F*DTKTA*KS*F*FA¥QRNLVK PRYGRKKKQKB *********** *********** GDF*****K** L8LINLAKVWSGADDVPAIRSNRTLRALRPLRKVSRWBGMK ******V*****LN*IKV*********************** ********I*A**A*I************************R «32) VSLINPVISLVGAGGIQAIKTNRTLRALRPLRANBRNDGIR *******'k*iA'k'k**************************** ***************************************** (G1) l7 O’Dowd and colleagues focused their study on exons a and i located along the same linker previously described to have PKA sites in mammalian SCs. Analysis of sodium current density in cultured embryonic neurons suggested that the presence of exon a was necessary for detection of sodium currents. They speculated that sodium current density is regulated by phosphorylation events because of a potential PKA site in exon 3 (O’Dowd et al., 1995). However, a later study investigating the role of exon a using Xenopus oocytes showed that the presence or absence of exon a did not alter expression of the sodium current (Warmke et al. 1997). To date, the functional consequence of exon a remains elusive. The functional importance of exon b has been examined in BgNav transcripts using the two-electrode voltage clamp system (Song et al., 2004). Deletion of exon b greatly enhanced the amplitude of sodium currents, whereas the inclusion of this optional exon dramatically reduced the amplitude of sodium currents in Xenopus oocytes. Interestingly, exon b contains a consensus sequence for phosphorylation that is hypothesized to serve as an on-off switch regulating neuronal excitability (Song et al., 2004). Another study of cockroach SCs looked into the functional consequences of mutually exclusive exons Gt, 62, and G3. Distinct gating and pharmacological properties were Observed for transcripts containing G1 and G2, however, transcripts containing G3 were non-functional due to the addition of a stop codon. Transcripts containing exon G1 activate and inactivate at more hyperpolarized membrane potentials in comparison to transcripts containing exon GZ. Channel sensivity to deltamethrin is greatly increased with the inclusion of 18 G1. Transcripts containing exon G2 have a 10-fold reduced sensitivity to deltamethrin than those containing Gt (Tan et al., 2002b). V.D RNA Editing Generates Diversity in Insect Sodium Channels RNA editing along with alternative splicing has been shown to generate molecular diversity in insect SCs. RNA editing is an important post- transcriptional modification that generates site-specific alterations of amino acids, in addition to insertions and deletions (references in Palladino et al., 2000a). A to l RNA editing is the result of deamination by the enzyme Adar (adenosine deaminase acting on RNA). This editase converts A to I through a process of hydrolytic deamination and has been reported to edit the SC in Drosophila melanogaster. Ten A to I RNA editing events have been identified in Drosophila para transcripts, and it appears that the frequency Of editing events dramatically increase with age (references in Palladino et al., 2000a). D. melanogaster mutants lacking Adar exhibit a variety of behavioral defects, providing profound evidence in support of the significance of A to I editing. Mutant larvae appeared normal in both locomotion and behavioral assays, while adults exhibit severe locomotion deficits and altered behavioral patterns. Abnormal behavioral phenotypes became increasingly acute with age. Tremors, strange body posture, slowness, difficulty mating, balance along with inefficient righting abilities, and circling behavior persistently worsened with age. Analysis of brain structure revealed brain lesions that intensified with age. Analysis of larval brains appeared normal while abundant lesions were present in adults. The extensive locomotive and behavioral defects seen in these mutants suggest 19 that ADAR editing events in the nervous system are necessary to perform normal function (Palladino et al., 2000b). In the German cockroach, B. gerrnanica, A to l and U to C RNA editing of BgNav has been reported. Similar to that of alternative splicing, tissue-specific and developmental-specific regulation of these RNA editing events has also been documented (Song et al., 2004). Two described U to C RNA editing events in BgNav undergo tissue-specific regulation; a leucine to proline change in DIIIS1 and a valine to alanine change in DIVS4. The tissues of ovary, gut, nerve cord, and leg were analyzed and both changes were found to only occur in the first two tissues listed (Song et al., 2004). Enzymes that catalyze or mediate U to C editing have not yet been identified. Also, an A to l editing event resulting in an isoleucine to methionine change in DIVS4 is regulated in a tissue specific manner. Transcripts of this A to I editing event were isolated from nerve cord and leg (Song et al., 2004). In addition to tissue-specific regulation, the A to l editing of DIVS4 is also developmentally regulated. Only the unedited transcript was found in the embryonic stage I while both edited and unedited were detected In two embryonic (II and III) and two immature (nymph l and II) stages. V.E Electrophysiological Analysis of RNA Edited Variants of BgNav RNA editing of BgNav along with alternative splicing produces 803 with distinct gating properties (Tan et al., 2002a; Liu et al., 2004, Song et al., 2004). Changes in current, activation, and inactivation are among some of the obvious results of RNA editing events. Examples include a U to C modification resulting in an F to S change at the C-terrninal domain that was found to impair 20 inactivation resulting in persistent current. It is speculated that the C-terrninal domain interacts with the IFM motif or docking receptor (discussed earlier) and modulates the kinetics of inactivation (Liu et al., 2004). An A to l editing event resulting in a lysine to arginine change in D182 causes a depolarizing shift in the voltage-dependence of activation by nearly seven millivolts (Song et al., 2004). To date, RNA editing events of the insect SC have only been documented and functionally characterized in the cockroach system. The U to C and A to l RNA editing events previously mentioned are the only two editing events that have been demonstrated to confer altered gating properties. RNA editing could be a major mechanism for increasing the diversity of sodium channel function. More research in this area needs to be done in order to characterize the importance of RNA editing and to understand the physiological roles it may play. VI. Pharmacology of Voltage-Gated Sodium Channels Sodium channels are integral transmembrane proteins responsible for the rising phase of action potentials (Catterall, 2000). Because of their fundamental role in neural excitability they are a target site of neurotoxins produced by animals and plants for defense and preying strategies as well as therapeutic drugs (Wang and Wang, 2003). Binding of toxins to 80$ change normal channel functional and can result in modifications ranging from blocking the channel pore to altering normal activation and inactivation processes (Wang and Wang, 2003). These toxins are grouped into nine types according to their distinct effects and receptor sites on the sodium channel (Table 4). The molecular determinants of receptor sites one through five, seven, and nine have been identified (Table 4). 21 Table 4: Neurotoxins that act on sod__i_gm channels Receptor Neurotoxins Physiological Location Site Effects 1 Tetrodotoxin Binds to extracellular 85-6 loop of saxitoxin pore; block Na” ion D1, D2, D3, D4 p-conotoxin entry 2 Batrachotoxin Persistent activation Segment 6 of veratridine D1, D2, D3, D4 grayanotoxin 3 a-scorption toxins Block Inactivation 83-4 loop of D4 sea anemone toxins 85-6 loop of D1, D4 4 B-scorpion toxins Prolong activation, 81-82 loop of D2 hyperpolarized shift 83-84 loop of DZ of activation 5 Brevetoxins Persistent Activation Segment 6 of D1 ciguatoxins Segment 5 of D4 6 6-conotoxins impair Inactivation Unknown 7 DDT, Block activation Segment 6 of D1, pyrethroids depolarized shift of D2, D3 resting potential 8 Conus stratius Prolong activation Unknown toxin 9 Local anesthetics, Block Inactivation, Segment 6 of D1 , prevent ion entry D2, D3 Modified from Wang and Wang (2003). The most relevant toxin this study is thalT, an alpha scorpion toxin extracted from Leiunrs quinquestn'atus hebraeus. Alpha scorpion toxins bind to receptor site three and remove fast inactivation resulting in prolonged sodium ion entry into the cell (Gordon et al., 1996). These neurotoxins are classified into 22 four groups based on pharmacological properties and structural similarities (Gordon et al., 1996). The ‘classical’ alpha scorpion toxins are highly active on mammalian 803 and weakly active on insect 803. The prototype for classical alpha scorpion toxins is AaH ll because of its high affinity towards mammalian brain synaptosomes (Gordon et al., 1996). Those toxins that are weakly toxic to both insects and mammals are represented by qu IV. These toxins competitively inhibit the binding of AaH II in mammalian systems as well as thalT in insect 808 (Gordon et al., 1996). “Alpha-like scorpion toxins’ are active on both mammalian and insect SCs. Interestingly though, alpha-like toxins Born Ill and Born IV compete with thodT for binding on insect 803, but not AaH ll of mammalian VGSCS. However, because of their “classical” inhibition of inactivation and competitive binding with thalT, they are termed ‘alpha-like’. Those toxins that are highly toxic to insects and weakly toxic to mammals are considered insect-selective. thodT is significantly more toxic to insect SCs than to mammalian 808 (Gordon et al., 1996). The a-scorpion toxins are small polypeptides of 63 to 70 amino acids cross-linked by four disulfide bridges (Cestele and Catterall, 2000). thaIT is greater than ten fold more active against insect 803 than mammalian SCs (Lee et al., 2000). Because the effect of thalT binding to insect SCs resembles the classical scorpion toxins suggests that the receptor site on insect 803 would be similar to mammalian 803. However, because thozIT exhibits selective tendencies towards insect sodium channels and not mammalian sodium channels indicates that there are unique structural features of the insect sodium channel that confers selective binding. It has been proposed that there are 23 distinct but overlapping receptor sites in insect and mammalian 80s for the binding of mammalian- and insect-selective toxins (Gordon et al.,1996). While photoaffinity labeling, site-directed mutagenesis studies, and site-directed antibodies against individual extracellular loops revealed Dl85-86, DlVS5—S6, and DIVS3-S4 are involved in alpha-scorpion toxin binding on mammalian sodium channels (references in Rogers et al., 1996), the critical amino acid residues involved in the binding to insect 803 have not been identified. 24 CHAPTER 2: Specific Aims, Rationale, and Significance Specific Aims: The Iong-terrn goals of this project are to understand the functional diversity of insect voltage-gated sodium channels (SCs); to elucidate the molecular interactions between 803 and neurotoxins particularly insecticides; and to understand the biological role of alternative splicing and RNA editing of insect 8C genes. The specific Objectives of my masters thesis project are: Objective 1. Expression and functional characterization of all 64 Para variants using a two-electrode voltage clamp /Xenopus oocyte expression system. Objective 2. Pharmacological analysis of the functional sodium channel variants using the a-scorpion toxin, thaIT. Rationale and Significance of Objectives: There exists one gene encoding 803 in insects. Transcripts of this gene undergo extensive alternative splicing and RNA editing events, generating functionally and phannacologically distinct sodium channels. A single composite full-length clone of para is currently available for functional, pharmacological, and molecular analysis. This clone was constructed by ligating two overlapping CDNA fragments (Warmke et al., 1997), and may not represent a true splice type or the variety of sodium channels in vivo. 25 Dr. Zhiqi Liu, a postdoctoral associate in Ke Dong’s laboratory, isolated sixty-four full-length clones of para by RT-PCR of the entire coding region. It is likely that these Clones represent a suite of in-vivo splice types. Sequence analysis of the full-length clones should reveal novel alternative splicing and RNA editing sites. Comparison of alternative exon use and RNA editing events between para and German cockroach SC genes would aid in the identification of common and distinct features of these events. Expression and functional analysis of the full-length CDNA clones in Xenopus oocytes is likely to reveal Para variants with unique gating properties similar to and differing from the properties previously identified in the German cockroach SCs (Tan et al., 2002; Song et al., 2004; Liu et al., 2004). This information will be the foundation for future experiments that will outline the importance of RNA editing and alternative splicing in vivo by taking advantage of the unique genetic tools available only in D. melanogaster. The SC is the primary target of various neurotoxins, including insecticides. Sodium channels exhibit distinct pharmacological profiles in both insect and mammals (Zhao et al., 2005; Song et al., 2006). This thesis will focus on the alpha scorpion toxin, thaIT for pharmacological analysis. thoIlT is an insect- specific scorpion toxin and little is known pertaining to the binding site of thole on insect 80s. The response of Para variants to thalT application will be determined. Identification of thaIT susceptible and resistant variants will provide valuable resources towards the analysis of molecular determinants that are critical for toxin binding and action. 26 CHAPTER 3: Functional and Pharmacological Characterization of Sixty- Four Para Sodium Channel Variants in Xenopus Oocytes ABSTRACT Extensive alternative splicing and RNA editing of the Drosophila melnaogaster voltage-gated sodium channel (SC) gene, para, has been documented. However, the functional consequences of these post- transcriptional modifications have not yet been investigated. In this study we report the isolation, expression, and functional characterization of sixty-four full- length para CDNA clones. The sixty-four clones were grouped into twenty-nine splice types based on alternative exon usage. Functional analysis of these variants revealed a broad spectrum in the voltage-dependence of activation with the voltages for half-maximal activation ranging from -6.8 mV to -46.8 mV. The voltage dependence of inactivation, however, was less variable. Sequence comparison of three functionally distinct variants uncovered three novel A to l editing sites. Furthermore, we revealed differential sensitivities of these variants to thalT, an a-scorpion toxin isolated from Leiurus quinquestn‘atus hebraeus. The observed functional and pharmacological diversity of para variants supports the notion that alternative splicing and RNA editing are the two major mechanisms for generating sodium channel diversity in insects. 27 INTRODUCTION Voltage-gated sodium channels (SCs) are integral transmembrane proteins responsible for the generation of action potentials across the membranes of excitable cells. The mammalian SC consists of a large, membrane bound, pore-forming, voltage-sensing a-subunit of approximately 260 kDa, along with several smaller auxiliary )9 subunits. The a-subunit is composed of four repeated homologous domains (I-lV), each containing six hydrophobic transmembrane segments (S1-86). Mammals have nine genes that encode for functionally and pharmacologically distinct SCs each with unique tissue, developmental, and cellular-specific expression patterns (see Chapter 1 for details; review by Catterall, 2000). Insect SCs, unlike mammalian counterparts, are encoded by a single gene. The Drosophila melanogaster SC, para, was isolated from temperature- sensitive paralytic mutants using P-element transposon tagging (Loughney et al., 1989). The complete Para amino acid sequence was deduced from six overtapping CDNA clones and was predicted to have extensive homology (greater than 58 percent for each four domains) to mammalian SCs (Loughney et al., 1989). Alternative splicing and RNA editing of para are predicted to be two major mechanisms for increasing molecular and functional diversity of insect SCs (Loughney et al., 1989; O’Dowd and Aldrich, 1988; O’Dowd, 1995; Thackeray and Ganetzky, 1994, 1995; Wannke et al., 1997; Palladino et al., 2000), but, the functional consequences of alternative exons and RNA editing sites have not been determined in para. However, the importance of alternative splicing and RNA editing of the German cockroach SC gene BgNav in creating insect SC 28 functional diversity has been recently documented (Tan et al., 20023; Liu et al., 2004; Song et al., 2004). Prior to our study a single variant of para has been functionally characterized in Xenopus oocytes (Warmke et al., 1997). However, this composite full-length CDNA clone may not represent a splicing type in vivo because it was constructed by ligating two overlapping partial CDNA clones. In this study we report the isolation, expression, functional and pharmacological characterization of sixty-four full-length para CDNA clones. MATERIALS AND METHODS Synthesis of first strand cDNAs Total RNA was isolated using lnvitrogen TRlzol Reagents (lnvitrogen, Rockville, MD). Isolation of mRNA was performed using Promega polyA+ RNA isolation kit (Promega, Madison, WI). First-strand CDNA was synthesized from mRNA using a Para-specific primer (D-3’-1; 5’ - TAC TCA TGC TAA TAC TCG CG -3’ ) based on the sequence of a region immediately down stream of the stop codon (Genbank accession M32078), and 8uper8cript ll RNase H-reverse transcriptase (lnvitrogen, Rockville, MD). First strand CDNA synthesis reaction conditions were: incubation at 42°C for 2 min. followed by a 60 min. incubation at 48°C. RNA was removed using RNaseH followed by a 20 min. incubation at 37°C. PCR and cloning of para full-length cDNAs To amplify the entire coding region of para (6 kb) by PCR, a fonIvard primer (D-para-Kpnl - 5’-CGG GGT ACC GCC ACC ATG GCA GAA GAT TCC GAC TCG-3’) and a reverse primer (D-para-Xba - 5’-GCT CTA GAT ACT GCT 29 AAT ACT CGC G—3’) were designed based on the 5’ and 3’ sequences of the para open reading frame. Xba and Kpnl restriction enzyme site sequences (underlined) were added to the forward and reverse primers, respectively, to facilitate cloning. A Kozak sequences (double underlined) was added to the primer D-para-Kpnl for high-efficiency translation. The reaction mixture of amplification (50 pl) contained 0.5 pl CDNA, 50 pmol each of primers D-para-Xba and D-para-Kpnl, 200 pM each dNTP, 1 U eLONGase (lnvitrogen), 1.5 mM MgCIz, and 1X PCR reaction buffer. PCR amplification was carried out as follows: 1 cycle of 94° C for 1 min.; 33 cycles of 94° C for 30 sec., 58° C for 30 sec., 68° C for 8 min.; one cycle of 68° C for 15 min. The PCR products were purified using QIAEX ll Gel Extraction Kit (QIAGEN Sciences, Maryland) and cloned in pGH19, an oocyte expression vector. Plasmids were transformed into competent bacterial cells, Stbl2 (lnvitrogen), and sixty-four positive colonies were isolated. Expression of para sodium channels in Xenopus Iaevis oocytes Oocytes were obtained surgically from oocyte-positive female Xenopus Iaevis (Nasco, Ft. Atkinson, WI). Females were anesthetized with Tricaine (3- aminobenzoic acid ethyl ester) at Zg/Iiter, and several lobes of the ovary were surgically removed. Follicle cells surrounding the oocytes were loosened by collagenase treatment (1 mg/ml Type IA collagenase, Sigma 00., 8t. Loise, MO) in calcium-free ND-96 medium (96 mM NaCl, 2mM KCI, 1 mM MQCIz, and 5 mM HEPES adjusted to a pH of 7.5) then the remaining follicle cells were removed with forceps. Isolated oocytes were incubated in N096 medium (96 mM NaCl, 2mM KCI, 1 mM MgCl2, and 5 mM HEPES, 1.8 mM CaClz, 50 [1ng gentamicin, 3O 5mM pyruvate, and 0.5 mM theophylline, adjusted to a pH of 7.5) (Goldin, 1992). Healthy oocytes stages V and VI were used for cRNA injection. Plasmid DNAs containing the para and TipE (auxiliary subunit of the para construct, refer to Chapter 1 for details) constructs were linearized by Notl. cRNAs were prepared by in vitro transcription with T7 polymerase (mMESSAGE mMACHINE kit, Ambion). TipE and para cRNAs were co-injected (0.25 ng — 10 ng) into Xenopus oocytes. TipE facilitates robust expression of para current in oocytes (see Chapter 1 for details). Oocytes with peak sodium current of 0.5 uA— 2 (IA were used for functional analysis using the two-electrode voltage clamp. Usually 0.5 pA- 2 uA of sodium current can be detected in the oocyte one to five days following cRNA injection. Typically, by the sixth day, the oocyte condition was too poor for recording currents. Electrophysiological recording and analysis Sodium currents were recorded using a standard two-electrode voltage clamp system (Kontis and Goldin, 1993). Voltage and current borosilicate glass electrodes were filled with 3 M KCI and manipulated for a resistance of less than 0.5 MI). Amplifier OC72SC (Warner Instrument Corp. Hamden CT), Digidata 1320A, and pCLAMP 8.2 software (both of Axon Instrument, Foster City, CA) were used for current measurement. All experiments were performed at room temperature (20 to 22°C). Capacitive transient and linear leak currents were corrected using the PIN subtraction technique. The voltage-dependence of conductance (G) was calculated by measuring peak currents at test potentials ranging from -80 to +65 mV in increments of 5 mV and dividing by (V-Vrev), where V is the test potential and Vrev is the reversal 31 potential for sodium. Reversal potentials were determined from the I-V curves and peak conductance values were normalized to the maximal peak conductance (Gmax). Peak conductance values were fitted with a Boltzmann equation of the form G=1-[1+exp(V-V1/2)/k]'1. Where V is the potential of the voltage pulse, Vuz is the half-maximal voltage for activation, and k is the slope factor. The voltage-dependence of inactivation was determined using 200ms inactivating prepulses ranging from -120 mV to -15 mV in 5 mV increments from a holding potential of -120 mV. Peak current amplitude during the test depolarization was normalized to the maximum current amplitude, and plotted as a function of the pre-pulse potential. The data were fit with a Boltzman equation of the form l=lmax*[1+(exp(V-V1/2)/k)]'1, where I."W is the maximal current, V is the voltage pulse potential, V1,; is the voltage at which half of the current is inactivated, and k is the slope factor. thaIT sensitivity assay thalT was generously provided by M. Gurevitz (Tel Aviv University, Israel). The method used for application of thalT in the recording system is identical to that described by Tan et al. (2002). Briefly, disposable recording chambers were made in Petri dishes with hot glue dams (1-1.5 ml volume). thalT was delivered to the oocyte recording chamber by hydrostatic force using glass capillary tubes. The toxin-induced effects were quantified by the ratio of the current amplitude 20 ms into depolarization (I20) over the peak current (lpeak). 32 RESULTS Molecular analysis of 64 full-length cDNA clones Prior to our study, eleven alternative exons were identified in the para transcript (Thackeray and Ganetzky, 1994; O’Dowd et al., 1995; Warmke et al., 1997; Lee et al., 2002). Seven of them are optional (a, b, e, f, h, i, j), and four are mutually exclusive (Old and UK; Fig.1A). In this study, Dr. Zhiqi Liu isolated sixty-four para full-length CDNA clones using RT-PCR. Determination of the alternative exon usage in these variants was done by using PCR or sequencing. Our analysis revealed twenty-nine splice types (Fig. 1B). The most abundant splice type is a, b, d, i, j, I. It is interesting that out of seven optional exons and four mutually exclusive exons that only twenty-nine splicing types were observed because there is a possibility of forming 20,160 splicing types. Most of the variants contain exons a, d, i, j, | suggesting that there is selection pressure for these exons. Functional expression of 64 variants In Xenopus oocytes All sixty-four of the para variants were co-injected with TipE into Xenopus oocytes for functional analysis using the two-electrode voltage clamp expression system. Twenty-six of the variants did not generate any sodium current even eight days after cRNA injection. Another six variants produced detectable sodium currents, but the currents were too small (less than 0.5 M) for functional analysis. Sodium currents from the remaining thirty-two variants were sufficient for functional analysis (0.5 IA -2 IA). We determined the voltage-dependence of activation and inactivation of these thirty-two variants. 33 As presented in Table 1, the voltages for half maximal activation (V1,2) varied greatly among the functional constructs. We divided the variants into three groups based on their voltage dependence of activation. Five variants, 50, 55, 48, 32, and 53, required more positive membrane potentials for activation with a Vm of -6.8 mV to 15.69 mV and were designated as high-voltage activated (HVA) variants. Nine variants, 49, 10, 45, 27, 62, 22, 2, 21, and 18, activated at more negative membrane potentials with V1/2 of -26 mV to -46 mV, and were termed low-voltage activated (LVA) variants. The majority of the variants had a VW value of -20 mV to -25 mV, and were labeled mid-voltage activated (MVA) variants. Voltages for half maximal inactivation (V10) were less variable for most of the variants with a range of -32 mV to -50 mV. Variant 18 was unique in that it inactivated at strongly hyperpolarized potentials (-65 mV) and also activated at extremely hyperpolarized potentials (-47 mV). Sequence analysis of three functionally distinct variants Three para variants were sequenced, each representing one of the three activation groups; LVA (variant 18), MVA (variant 13), and HVA (variant 50). Each variant contained unique alternative splicing and RNA editing sites. Variant 13 belonged to splice type fourteen (c, i, j, I), variant 18 belonged to splice type seven (a, c, j, l), and variant 50 belonged to splice type nineteen (d, e, f, i, j, k). 34 [1A 3 15543332222221111111111111111 Vanant Number k W WWW W EE k IEEEEEEEEEEEEEEEEEEMEEEEEE EQI ELI“ 6 “mm m H mm I. e m m a e as WWW SSW .emm a IWII E lella I... W E Ewmiid It bEEEE E E E E EE W E E E b an an awn an an :::::m l/k / / WA d d .m. 0.123456789012345 T1234567891111111111222222 U LI] T-U ta . . mm a v/.- -mmmmmm mmmmm mmmmmmmm mmm Splicing 26 A. A topological view of the sodium channel with positions of alternative exons indicated by colored boxes (A). 3. Schematic presentation of alternative exon usage in each splice type. 35 Identification of 29 splice types. Figure 1. Table 1. Activation and inactivation properties of 32 functional variants. Variant Activation Inactivation V1,2(mV) k (mV) n Vuz (mV) k (mV) n 50 -6.80 11.74 5.20 1 1.10 12 60.83 1 1.69 6.11 1 0.68 18 55 -9.94 1 3.72 9.39 1 4.56 6 —50.19 1 2.95 5.42 1 0.85 10 48 41.42 1 2.34 7.33 1 0.79 14 34.90 1 3.86 5.47 1 1.06 14 32 -14.78 1 3.24 7.27 1 0.96 21 42.32 1 5.22 7.31 1 6.20 39 53 ~15.69 1 2.05 8.16 1 1.06 10 -45.46 1 1.30 7.46 1 0.63 12 T 49.37 1 1.81 6.88 1 1.20 20 -39.81 1 2.88 6.06 1 1.50 37 38 -20.49 1 4.63 7.94 1 1.37 8 -46.72 1 0.46 4.84 1 0.55 7 17 -2057 1 2.72 6.78 1 1.31 13 46.38 1 2.69 4.98 1 0.43 18 58 -21.60 1 2.68 6.33 1 1.22 23 40.99 1 2.02 5.28 1 0.65 24 24 -21.60 1 2.07 6.90 1 1.19 21 43.95 1 2.14 5.36 1 0.72 25 7 -22.40 1 2.48 7.44 1 1.19 11 47.41 1 2.78 5.12 1 0.70 15 19 -2229 1 3.07 8.68 1 1.57 19 45.39 1 2.27 5.44 1 0.79 25 11 -22.71 1 3.75 8.11 11.83 19 47.32 1 2.12 5.55 1 0.42 23 13 -2274 1 3.33 3.46 1 0.64 11 -32.07 1 2.17 3.85 1 0.49 17 39 -23.26 1 4.58 5.49 1 1.54 19 -39.87 1 2.90 5.29 1 1.10 13 41 -23.70 1 3.33 4.50 1 0.83 24 40.09 1 2.65 4.54 1 1.01 24 24 -23.77 1 3.72 6.69 1 0.85 14 -48.31 1 2.30 5.32 1 0.48 15 31 -2403 1 2.98 5.15 1 0.80 12 45.06 1 1.97 4.83 1 0.74 22 47 -25.34 1 4.34 4.70 1 1.25 20 42.87 1 2.59 4.82 1 0.33 20 57 -25.66 1 2.23 5.80 1 1.23 3 46.07 1 3.08 4.87 1 0.36 15 56 -25.72 1 4.06 5.94 1 1.39 20 -35.05 1 3.09 4.94 1 0.42 21 8 -25.75 1 1.95 6.38 1 1.58 8 47.11 1 2.20 4.92 1 0.33 11 44 -2599 1 4.49 3.00 1 1.30 24 -39.21 1 1.46 4.18 1 0.44 24 _49’— -26.25 1 2.66 5.08 1 1.01 17 43.51 1 1.59 4.85 1 0.30 19 10 -26.33 1 5.77 3.95 1 1.19 18 -38.95 1 1.64 4.84 1 0.51 19 45 -27.13 1 3.80 5.81 1 1.04 24 45.15 1 2.64 5.35 1 0.21 24 27 -27.16 1 3.92 4.91 1 1.13 21 41.86 1 3.18 5.02 1 0.30 22 62 -29.08 1 3.58 3.68 1 0.87 17 42.88 1 1.82 4.49 1 0.37 18 22 -29.53 1 2.48 10.36 1 2.07 13 49.19 1 1.95 6.48 11.19 13 2 -29.67 1 7.27 4.35 1 1.19 33 42.01 1 2.45 5.05 1 0.52 37 21 -31.66 1 3.69 4.32 1 0.90 20 43.66 1 2.07 4.55 1 0.29 25 18 -46.56 1 4.55 6.17 1 1.06 17 -65.42 1 2.42 4.43 1 0.46 24 Sequence comparison of variants 13, 18, and 50 with the GenBank Para sequence (accession number M32078) revealed fifteen, twelve, and eight amino acid changes, respectively (Figure 2 and Table 2). In an earlier study, ten A to I 36 editing sites were identified in para (Palladino et al., 2000a). Five of the observed Changes from this study match five of the previously reported A to l editing sites (Palladino et al., 2000). Because PCR mistakes are unlikely to occur in multiple variants, we have identified and additional three A to I editing sites (observed in multiple variants; Table 2A). Additionally, two nucleic acid Changes were observed in multiple variants that resulted in a guanine to cytosine change. However, whether these are the result of alternative splicing or a novel type of RNA editing remains to be determined. Nucleic acid changes that occur in a single variant may be the result of RNA editing, alternative splicing, or a mistake made during PCR. Fifteen distinct nucleic acid changes were observed in variants 13, 18, and 50 (Table 23). One of these changes, N1822S, corresponds to a previously recognized A to l editing site (Palladino et al., 2000a). The remaining fourteen changes need to be further investigated to determine the foundation of change. 37 E2990 D1300N M1376V T1532A I1661V |1691V N18228 Variant 13 R520 UU Q471R R12960 L1363F N1565K |167O I1679V E2990 K303R W319R I1661V |1691V A1731V An U L1363F |1670M H1676 E299Q 89 2P D1300N I1661V |1691V N18228 , AA X Variant 50 R520 UV U 'L' R12960 L1363F S1587N |1670 I1679V Figure 2. Amino acid changes in variants 13, 18, and 50. The secondary structure of variants 13, 18, and 50 along with the position of amino acid change compared to the GenBank sequence (M32078). For example, R520 indicates R in the Genebank sequence and Q in variants at the amino acid position 52. Amino acid changes occur in more than one variant are indicated by the darkened circles. Triangles represent variant-specific amino acid changes. 38 Table 2. Amino acid changes observed in variants 13, 18, and 50. Variant Variant Variant Nucleo- Location RNA Report- 13 18 50 tide Editing ed Change Type A to l A R 52 Q R 52 Q R 52 Q g-a N.T. A to l * E 299 Q E 299 Q E 299 Q J—C l85-l86 ? Q 471 R Q 471 R a-l I-ll Atol R 1296 Q R1296 Q g-a ll-lll A to l D 1300 N D1300 N J-a IIIS1 A to l L 1363 F L 1363 F g—C lll82-lll83 ? l 1661 V I 1661 V a-g IVS1-IVS2 Atol l 1670 M l 1670 M a-g IVS2 A to l I 1691 V l 1691 V a-g IVS3 Atol B I 260 T a-g 184-185 K303R a-g l84-l85 W319 t-c |85-IS6 S 922 P t-c IIS4 L1363F g-c I|82-IIS3 M 1376 V ag Ill83 Q 1450 R Q llIS5 T 1532 A a-g IllS5-III86 N 1562 K t:g Ill-IV S 1587 N g-a Ill-IV H1676R a-g IV82-IVS3 I 1679 V a;g IVS3 A 1731 V c-t IV84 N 1822 S a-g lVS5-IVS6 * l 1850 M EL IVS6 Note: N.T. represents the N-terrninal domain. Question marks (7) indicate that the mechanism underlying the amino-acid changes are unknown. Asterisks (*) represent the previously reported A to I editing sites in para (Palladino et al., 2000). Examination of the sensitivity of Para variants to thalT thalT, an alpha scorpion toxin isolated from Leiurus quinquestn'atus hebraeus, binds to receptor site three (refer to Chapter 1) and removes fast inactivation resulting in prolonged action potential propagation (Chen et al., 39 2000). thadT is polypeptide of 66 amino acids that is ten fold more selective to insect 803 than to mammalian SCs (Lee et al., 2000). To examine the sensitivities of the functional variants of para to 1 nM thalT, we determined the ratio of Izo/lpeak before and after toxin application (Figure 3; Table 3). Although inactivation was affected by 1 nM thalT for all variants, some variants are more sensitive than others. For example, variant 21 is most resistant to thalT compared to all other variants with only 29% of inactivating current inhibited by 1 nM thalT. Variants 8 and 58 are among the most sensitive ones with inactivation being completely abolished by toxin application. To quantitatively compare the effect of thalT between variants 8 and 21. dose-response curves were created (Fig. 4). While the E050 of variant 8 could not be generated due to an inadequate number of data points, Figure 4 indicates that variant 8 is apparently more sensitive than variant 21 (E050 of 2.58 nM) to thalT. Table 3. Effects of1 nM thalT on para variants Variant '20/lpeak (%) 1 SD n Variant I20/Ipeak (%):|; SD n 21 29113 12 19 62:18 3 62 34 :I: 13 4 55 66 :I: 22 3 2 58133 27 13 78136 13 10 34 :I: 13 4 11 79 :l: 16 3 22 34 :I: 13 4 8 80 :l: 25 18 47 29 :t 13 12 17 81 :l: 9 4 56 34 :l: 13 4 50 85 i: 37 2 39 58 :I: 33 27 7 86 :I: 12 2 48 60 j: 15 6 31 88 1 18 61 :I: 17 6 44 88 1 25 56 :I: 14 5 27 93 1 45 58 1 24 94 :t 1 4 49 58 1 52 105 :I: 7 8 41 59 :I: 0 2 32 105 1 57 62 1: 27 4 58 105 :t 9 2 40 1 nM thalT I I -1-0011A 1.84 “A 1 I Figure 3. Peak sodium current traces of variant #8 before and after application of 1 nM thalT. Peak current was measured by depolarizing the oocytes membrane to - 10 mV for 20 ms from a holding potential of -120 mV. Peak current was determined before the application of 1 nM thalT and ten minutes after the application of 1 nM thaIT. 140 - “2% 20 1 " a. 1 ‘g 004 I f II 1' =- I C 80 I -- 31/ .2 f/ = 1 /_/ .0 e E 60 ‘ '//’/ s _. T 5 40 -b ///-l . var'ant # 21 2 ‘ --”’ 0 20 1 I-‘/” I varlant # 8 0- ,.~ . 0 V V ' V v '1 I I “"10 thalT Concentration (nM) Figure 4. Dose response curves of variants 8 and 21. Dose-response curves were fit with a Hill equation of the form y = Vmx [x"/(k" + x")]. Variant 21 has an E050 of 2.58 nM. More data points were needed to determine the EC50 of variant 8. 41 Variants 8 and 21 were selected for sequence analysis because of their differential sensitivities to thodT. The results of this analysis are presented in Figure 5 and Table 4. Eight of the observed changes correspond to the previously recognized RNA editing sites. Three of the changes are identical to the earlier mentioned possible RNA editing sites and five are novel. Variant 8 has one novel amino acid change (K1455R) and variant 21 has two novel amino acid changes (|1519V and H1772Y) that are located on the extracellular loops of the pore and that may interact with the binding of thodT. 52990 K1455R I1661V |1691V N1822S Variant 8 R5 L1363F |1670 I1679V £2990 D1300N I1519V I1661Vl1691V N18228 A An Variant s 21 R520 UU U |260VH487N DB 66 L1363F |1670M I1679V H1772 R12960 Figure 5. Amino acid changes in variants 8 and 21 compared to GenBank # M32078. This figure depicts the topology of variants 8 and 21. Circles indicate common changes. Triangles represent unique changes in the transmembrane regions. and stars indicate unique changes in the extracellular portion. 42 Table 4. Sequence comparison of variants 8 and 21 Variant 8 Variant 21 Nucleotide Location Change R 52 Q1 R 52 Q“ g-a N.T. I 260 v2 a-g lS4-SS E 299 Q‘ E 299 Q1 g-c lSS-ISG H 487 N3 c-a |SG-llS1 D 656 G3 a-g ISG-lIS1 R 1296 Q" j-a IISG-IIIS1 D 1300 N1 g-a IIIS1 L 1363 F“ L 1363 F“ g-c llISZ-lllS3 K 1455 R4 a-g lllS5-SG I1519V" a-g lllS5-36 I 1661 v1 I 1661 VT a-g Ivs1-Ivs2 l 1670 MI I 1670 M‘ a-g IVS2 I 1679 v2 I 1679 v2 a3 IVS3 I 1691 v1 I 1691 VI a-g IVS3 H 1772 Y4 c-t Ivss-se N 1822 s2 N 1822 $2 a-g IVS5-IVS6 Note: 1. Amino acid changes that correspond to previously recognized RNA editing sites of variants 13, 18, and 55. 2. Amino acid changes that correspond to possible RNA editing sites of variants 13, 18, and 55. 3. Unique editing sites. 4. Unique editing sites that are situated on the extracellular portion of the pore and that may interact with the binding of thalT. DISCUSSION: Functional versus non-functional Para variants in oocytes In this study, we co-expressed the sixty-four cDNA clones of para with TipE to achieve robust expression of sodium currents as previously demonstrated by Warmke et al. (1997). Half of the total variants, thirty-two, had sodium currents large enough for functional analysis (1-2 M). An additional six variants were functional, however, did not produce sufficient sodium currents for functional analysis (<1 IA). The remaining twenty-six variants did not generate any detectable sodium current (0 [.LA). Recently, four TipE-homologous genes (T EH1-4) were identified from D. melanogaster (Derst et al., 2006). The co-expression of three of these genes 43 (THE 1-3) with para (the para construct was obtained from Dr. J.C. Cohen; refer to Warmke et al., 1997) in Xenopus oocytes also elevated sodium current expression. It would be of interest to determine whether TEH1-4 can boost the low level expression of the six variants in our study. RT-PCR analysis indicated that the expression patterns of TEH1 and TEH4 are under tissue-specific regulation (Derst et al., 2006). Distinct tissue distribution patterns have also been observed for alternative splicing and RNA editing of BgNav variants (Tan et al., 2002; Song et al., 2002). The possible tissue-specific co-expression of para variants with TipE or THE1-4 in vivo remains to be investigated. The presence of exon b in para and the ortholog B in BgNav hampers the expression of BgNav and Para channels expressed in oocytes (Song et al., 2004; Song, Personal Communication). Interestingly, four of the six variants with low levels of sodium current expression contain exon B/b, suggesting that this exon could be the cause of poor expression. However, the majority of our functional variants also contain this exon (eighteen out of thirty-two), suggesting that the expression of sodium current is not solely regulated by exon B/b. At present, we do not know why the twenty-six variants are not functional in Xenopus oocytes. Sequencing of the twenty-six nonfunctional variants will reveal whether these variants contain stop codons in the open reading frame thereby generating non-functional channels. It is also possible that these variants require one of the other auxiliary subunits, such as TEH1-4, to facilitate expression, or that unidentified sequence features in these variants down- regulate sodium current expression. Identification of twenty-nine Para splice types Theoretically 20,160 splicing types could be formed out of the seven optional exons and four mutually exclusive exons. Interestingly though, we only isolated twenty-nine splice types. Although we may have failed to isolate some low frequency splicing types, we believe these twenty-nine represent the majority of transcripts in vivo. A previous study investigating exon use of BgNav variants identified 20 splice types from a total of 69 variants (Song et al., 2004). These results suggest that alternative exon usage is regulated, and not formed by random combinations. Since molecular analysis of a large pool of insect SC full-length cDNA clones has been conducted in both para (this study) and BgNav (Song et al., 2004), a direct comparison of splice types can be made between the two. The most abundant splice type of para (twenty percent) is a, b, d, iI i, kI and I, while the most abundant splice type (fifty-five percent) of BgNav is a. f. i. k and l. lntriguingly, the most abundant para splice type was not isolated at all from BgNav variants (Song et al., 2004). This discrepancy is due in part to the differences in the frequency of exon usage of several alternative exons (described above). Frequency of alternative splicing of insect sodium channels Previous studies show that alternative splicing sites of SC genes are extremely conserved among D. melanogaster, D. virilis, M. domestica, and B. gerrnanica (See Chapter 1). However, the frequencies of alternative exon use are less conserved (Table 5). In general, similar exon frequencies can be observed for closely related species, D. melanogaster and D. virilis, and even for 45 M. domestica SCs. The frequency of exon use between B. gennanica, D. melanogaster, D, virilis and M. domestica SCs, however, are more divergent. For example, exon a is frequently used in D. melangoaster, D. virilis, M. domestica and B. gennanica SC variants while exon e is not. Exons b, d, and i are expressed in greater than sixty percent of the para (both D. melanogaster and D. vin'Iis) and M. domestica variants. However, exons b and i are not commonly used in BgNav transcripts and use of exon d has not yet been identified in BgNav. Exons c and fare not frequently used in D. melangoaster, D. vin'lis, or M. domestica SC transcripts, but are present in over ninety-eight percent of BgNav transcripts. Exon h is used in low frequencies in para, both of D. melanogaster and D. virilis, and in BgNav, however, this exon is present in all transcripts of Vssc1. The frequency of expression of mutually exclusive exon j is high for para and BgNav transcripts (eighty-nine and seventy nine percent respectively) but low for Vssc1 (D. virilis has not been analyzed for the frequency of exon j) identification of Three Novel RNA Editing Sites Prior to this study, ten A-to-l editing sites were identified in para (Palladino et al., 2000), and two were reported in BgNav variants (Liu et al., 2004; Song et al., 2004). Five of the ten A-to-l editing sites identified in para were also identified in the three sequenced variants (13, 18, and 50). Additionally, we identified three novel A-to-I editing events, resulting in three amino acid changes, two l-to-V and one l-to-M change. Potentially there are more RNA editing sites in these three sequenced variants, however, alternative splicing or a mistake made during PCR can not be ruled out. Investigations into the additional changes 46 observed revealed three possible U-to-C editing. In a previous study three U-to- C editing sites were reported in BgNav variants (Liu et al., 2004; Song et al., 2004). Additionally, there are nine possible A-to-l editing sites and two novel changes that do not conform to either A-to-l or U-to-C editing. Table 5. Frequencies of Exon Usage in Para, VSSC1, BgNav Exon Para1 Para2 D. virilisr VSSC1‘ BgNav!s A 71.9 57.3 54.4 91-98 98.5 B 60.9 ~80 ~70 100 17.4 C 29.7 <30 ~20 ~0 100 D 70.3 >70 ~80 ~100 0 E 18.8 ~25 ~25 <10 17.2 F 9.4 6.7 ~10 ~10 98.5 H 3.1 21 100 0 I 85.9 ~85 81 100 5.8 J 89.1 <25 79.7 K 3.1 78-100 5.8 L 96.9 <25 92.7 1 . This study 2. O'Dowd et al., 1995; Thackeray and Ganetzky 1994, 3. Thackeray and Ganetzky, 1995 4. Lee et al., 2002;. 5. Song et al., 2004 Estimates are indicated by ‘~' marks, ‘>’ means greater-than, and ‘<’ means less than. It is worthwhile mentioning here that a previously identified U-to-C editing event in both para and BgNav was not detected in any of these sixty-four variants. This U-to-C editing event results in persistent current (Liu et al., 2004) and none of our thirty-two functional variants exhibited persistent currents. Obviously we have missed this variant in our cloning due to the low frequency of this editing event. For the same reason, we may have missed the isolation of other variants which have low editing frequency and play critical roles in regulating inactivation kinetics in vivo. 47 Identification of High-Voltage Activated and Low-Voltage Activated Para Variants The functional analysis of our sixty-four para cDNA clones revealed a broad spectrum in the voltage-dependences of activation among the variants. In an earlier study performed on BgNav variants a broad range in the voltage- dependency of activation was also identified (Song et al., 2004). The range of the voltage-dependences of activation for the BgNav variants was from -24.9 mV to -43.9 mV (Song et al., 2004). In this study we, for the first time, identified several LVA and HVA sodium channels. Variant 18 is an LVA channel with a voltage—dependence of activation of -46 mV. Variants 50 and 55 are extreme HVA channels with voltage dependences less than -10 mV. An additional sixteen of our variants activate at potentials less than that previously reported for BgNav variants (less than -24.9 mV). The molecular basis of the LVA and HVA gating characteristics remained to be identified. Mammalian SCs exhibit distinct voltage dependences of activation (Chapter 1 — Table 1). In general, those isoforms, that are predominantly located in the central nervous system (CNS; Nav 1.1, 2, 3, and 6) activate at more depolarized (i.e., positive) potentials than those predominantly located in the peripheral nervous system (PNS; Nav 1.7, 8, and 9). The cardiac sodium channel, Nav 1.5, and isoform Nav 1.9 activate at more hyperpolarized (i.e., negative) potentials than the remaining mammalian sodium channels (Chapter 1 - Table 2). The half-maximal voltages for activation of Nav1.5 (heart SC) are -45 or -46 mV and Nav1.9 (DRG neurons) are -47 or -54 mV, similar to that of the Para variant 18. 48 The terms, HVA and LVA, are well-established to describe distinct calcium channel families (review by Yunker, 2003). LVA calcium channels, such as Cav3.1, 2, and 3, activate at hyperpolarized potentials near resting membrane potential, while HVA calcium channels, such as Cav1.1, 2, 3, and 4, require strong depolarization above resting membrane potential in order to activate the channel (review by Catterall et al., 2005). LVA calcium channels are responsible for the repetitive firing of action potentials in pacemaker cells while HVA channels are involved in excitation-contraction coupling, hormone release, and neurotransmitter release (review by Catterall et al., 2005). Whether these Para LVA channels play a similar role and have similar tissue distributions as mammalian Nav 1.5 and LVA calcium channels remains to be determined. Both Nav1.5 and LVA calcium channels are localized in the heart muscle and both facilitate spontaneous firing. While we do not know the tissue distribution pattern of variant 18, it may be expressed in tissues and cells with pace-making activities. Studies of cockroach terminal abdominal ganglia DUM neurons characterized pacemaking activities as the result of LVA sodium channels (Lapied et al., 1998). These LVA sodium channels activated at potentials of -50 mV, which corresponds to the resting membrane potential of DUM neurons in cockroach terminal abdominal ganglia. It was hypothesized in this study that background channels are essential in the instigation and continuation of spontaneous firing because activation of these channels depolarize membrane potentials which in turn activate additional channels and result in action potential propagation. It would be of interest to determine the 49 tissue distribution of variant 18, and determine if this variant is responsible for the pace-making activities observed in the DUM neurons. Identification of an thalT-resistant Para variant The specific amino acids of the SV that are critical for thorlT binding remain elusive. Identification of an thorlT-resistant variant, #21, from our survey of thalT sensitivity of sixty-four Para variants provided valuable information for further elucidating the molecular determinants of the thalT binding site on the sodium channel. Photoaffinity labeling, site-directed mutagenesis studies, and site-directed antibodies against individual extracellular loops revealed DlSS—SG, DlVS5-86, and DlVSS-S4 are involved in a-scorpion toxin binding on mammalian sodium channels (references in Rogers et al., 1996). We found a single novel amino acid change in variant 8 that is located in the extracellular loop connecting DlllSS to Dll|86 that may confer the observed sensitivity of this variant to thalT binding (K1455R). The insensitive variant 21 has two novel amino acid changes (|1519V and Y1772H) that are located in the extracellular loops connecting $5 to $6 of Dlll and DIV that may interact with thalT and inhibit binding of the toxin to the channel resulting in a resistant variant. Future mutagenesis studies are needed in order to verify the involvement of these amino acids in thalT binding. 50 LITERATURE SITED Agnew WS, Moore AC, Levinson SR, Raftery MA. 1980. Identification of a large molecular weight peptide associated with a tetrodotoxin binding protein from the electroplax of Electrophorus electricus. Biochem Biophys Res Commun 92:860-6 Beneski DA, Catterall WA. 1980. Covalent labeling of protein components of the sodium channel with a photoactivable derivative of scorpion toxin. Proc Natl Acad Sci U S A 77:639-43 Castella C, Amichot M, Berge JB, Pauron D. 2001. DSC1 channels are expressed in both the central and the peripheral nervous system of adult Drosophila melanogaster. Invert Neurosci 4:85-94 Catterall WA. 1986. Molecular properties of voltage-sensitive sodium channels. Annu Rev Biochem 55:953-85 Catterall WA. 2000. From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels. Neuron 26:13-25 Catterall WA, Goldin AL, Waxman SC. 2003. lntemational Union of Pharmacology. XXXIX. Compendium of voltage-gated ion channels: sodium channels. Pharmacol Rev 55:575-8 Catterall WA, Goldin AL, Waxman SC. 2005. lntemational Union of Pharmacology. XLVIl. Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol Rev 572397-409 Cestele S, Catterall WA. 2000. Molecular mechanisms of neurotoxin action on voltage-gated sodium channels. Biochimie 82:883-92 Chen H, Gordon D, Heinemann SH. 2000. Modulation of cloned skeletal muscle sodium channels by the scorpion toxins th ll, th III, and th alphalT. Pflugers Arch 439:423-32 Claes L, Del-Favero J, Ceulemans B, Lagae L, Van Broeckhoven C, De Jonghe P. 2001. De novo mutations in the sodium-channel gene SCN1A cause severe myoclonic epilepsy of infancy. Am J Hum Genet 68:1327-32 Derst C, Walther C, Veh RW, Wicher D, Heinemann SH. 2006. Four novel sequences in Drosophila melanogaster homologous to the auxiliary Para sodium channel. Biochem Biophys Res Commun 339:938-48 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 51 Eaholtz G, Colvin A, Leonard D, Taylor C, Catterall WA. 1999. Block of brain sodium channels by peptide mimetics of the isoleucine, phenylalanine, and methionine (IFM) motif from the inactivation gate. J Gen Physiol 1 132279-94 Eaholtz G, Scheuer T, Catterall WA. 1994. Restoration of inactivation and block of open sodium channels by an inactivation gate peptide. Neuron 12:1041- 8 Germeraad S, O'Dowd D, Aldrich RW. 1992. Functional assay of a putative Drosophila sodium channel gene in homozygous deficiency neurons. J Neurogenet 8:1 -1 6 Goldin AL. 1999. Diversity of mammalian voltage-gated sodium channels. Ann N YAcad Sci 868:38-50 Goldin AL, Barchi RL, Caldwell JH, Hofmann F, Howe JR, et al. 2000. Nomenclature of voltage-gated sodium channels. Neuron 28:365—8 Goldin AL. 2002. Evolution of voltage-gated Na+ channels. J Exp Biol 205:575-84 Goldin AL. 2003. Mechanisms of sodium channel inactivation. Curr Opin Neurobiol 13:284-90 Gordon D, Martin-Eauclair MF. 1996. Scorpion toxins affecting sodium current inactivation bind to distinct homologous receptor sites on rat brain and insect sodium channelsd. J Biol Chem 271 :8034-45 Guy HR, Seetharamulu P. 1986. Molecular model of the action potential sodium channel. Proc Natl Acad Sci U S A 832508-12 Heinemann SH, Terlau H, Stuhmer W, lmoto K, Numa S. 1992. Calcium channel characteristics conferred on the sodium channelby single mutations. Nature 356:441-3 Hille B. 1984. Ionic Channels of Excitable membranes Sinauer Associates, Inc, Sunderland, MA Hodgkin AL, Huxley AF. 19523. The dual effect of membrane potential on sodium conductance in the giant axon of Loligo. J Physiol 116:497-506 Hodgkin AL, Huxley AF. 1952b. A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol 1 17:500-44 52 Hong CS, Ganetzky B. 1994. Spatial and temporal expression patterns of two sodium channel genes in Drosophila. J Neurosci 14:5160-9 lsom LL. 2001. Sodium channel beta subunits: anything but auxiliary. Neuroscientist 7:42-54 Kelly P, Yang WS, Costigan D, Farrell MA, Murphy S, Hardiman O. 1997. Paramyotonia congenita and hyperkalemic periodic paralysis associated with a Met 1592 Val substitution in the skeletal muscle sodium channel alpha subunitua large kindred with a novel phenotype. Neuromuscul Disord 72105-11 Kontis KJ, Rounaghi A, Goldin AL. 1997. Sodium channel activation gating is affected by substitutions of voltage sensor positive charges in all four domains. J Gen Physiol 110:391-401 Kontis KJ, Goldin AL. 1993. Site-directed mutagenesis of the putative pore region of the rat lIA sodium channel. Mol Pharmacol. 43:635-44 Lee SH, Ingles PJ, Knipple DC, Soderlund DM. 2002. Developmental regulation of alternative exon usage in the house fly Vssc1 sodium channel gene. Invert Neurosci 4:125-33 Lee D, Gurevitz M, Adams ME. 2000. Modification of Synaptic Transmission and Sodium Channel Inactivation by the Insect-Selective Scorpion Toxin thalT. J. Neurophysiol83z1181-7 Liu Z, Chung l, Dong K. 2001. Alternative splicing of the BSC1 gene generates tissue-specific isoforms in the German cockroach. Insect Biochem Mol Biol 31 2703-13 Liu Z, Song W, Dong K. 2004. Persistent tetrodotoxin-sensitive sodium current resulting from U-to-C RNA editing of an insect sodium channel. Proc Natl Acad Sci U S A 101 :1 1862-7 Loughney K, Kreber R, Ganetzky B. 1989. Molecular analysis of the para locus, a sodium channel gene in Drosophila. Cell 58:1143-54 McPhee JC, Ragsdale DS, Scheuer T, Catterall WA. 1998. A critical role for the 84-85 intracellular loop in domain IV of the sodium channel alpha-subunit in fast inactivation. J Biol Chem 273:1121-9 Noda M, lkeda T, Suzuki H, Takeshima H, Takahashi T, et al. 1986. Expression of functional sodium channels from cloned cDNA. Nature 322:826-8 Noda M, Shimizu S, Tanabe T, Takai T, Kayano T, et al. 1984. Primary structure of Electrophorus electricus sodium channel deduced from cDNA sequence. Nature 312:121-7 53 O'Dowd DK, Aldrich RW. 1988. Voltage-clamp analysis of sodium channels in wild-type and mutant Drosophila neurons. J Neurosci 8:3633-43 O'Dowd DK, Germeraad SE, Aldrich RW. 1989. Alterations in the expression and gating of Drosophila sodium channels by mutations of the para gene Neuron. 421301-11 O'Dowd DK, Gee JR, Smith MA. 1995. Sodium current density correlates with expression of specific alternatively spliced sodium channel mRNAs in single neurons. J Neurosci 15:4005-12 Palladino MJ, Keegan LP, O'Connell MA, Reenan RA. 2000a. dADAR, a Drosophila double-stranded RNA-specific adenosine deaminase is highly developmentally regulated and is itself a target for RNA editing. Rna 6:1 004-18 Palladino MJ, Keegan LP, O'Connell MA, Reenan RA. 2000b. A-to-l pre-mRNA editing in Drosophila is primarily involved in adult nervous system function and integrity. Cell 102:437-49 Patton DE, West JW, Catterall WA, Goldin AL. 1992. Amino acid residues required for fast Na(+)—channel inactivation: charge neutralizations and deletions in the lII-IV linker. Proc Natl Acad Sci U S A 89:10905-9 Ramaswami M, Tanouye MA. 1989. Two sodium-channel genes in Drosophila: implications for channel diversity. Proc Natl Acad Sci U S A 86:2079—82 Rogers JC, Qu Y, Tanada TN, Scheuer T, Catterall WA. Molecular determinants of high affinity binding of alpha-scorpion toxin and sea anemone toxin in the S3-S4 extracellular loop in domain IV of the Na+ channel alpha subunit. J Biol Chem 271:15950-62 Rohl CA, Boeckman FA, Baker C, Scheuer T, Catterall WA, Klevit RE. 1999. Solution structure of the sodium channel inactivation gate. Biochemistry 38:855—61 Saito M, Wu CF. 1991. Expression of ion channels and mutational effects in giant Drosophila neurons differentiated from cell division-arrested embryonic neuroblast. J. Neurosci 722135-50 Salkoff L, Butler A, Wei A, Scavarda N, Giffen K, et al. 1987. Genomic organization and deduced amino acid sequence of a putative sodium channel gene in Drosophila. Science 237:744-9 Siddiqi O, Benzer S. 1976. Neurophysiological defects in temperature-sensitive paralytic mutants of Drosophila melanogaster. Proc Natl Acad Sci U S A 73:3253-7 54 Silver K, Soderlund DM. 2005. State-dependent block of rat Nav1.4 sodium channels expressed in xenopus oocytes by pyrazoline-type insecticides. Neurotoxicology 26:397-406 Soderlund DM. 2005. Sodium channels. Comp MoI Insect Sci 5:1-24 Song W, Liu Z, Tan J, Nomura Y, Dong K. 2004. RNA editing generates tissue- specific sodium channels with distinct gating properties. J Biol Chem 279:32554-61 Song W, Liu Z, Dong K. 2006. Molecular basis of differential senstivity of insect sodium channels to DCJW, a bioactive metabolite of the oxadiazine insecticide indoxacarb. Neurotoxicology 2:237-44 Striessnig J, Glossmann H, Catterall WA. 1990. Identification of a phenylalkylamine binding region within the alpha 1 subunit of skeletal muscle Ca2+ channels. Proc Natl Acad Sci U S A 87:9108-12 Stuhmer W, Conti F, Suzuki H, Wang XD, Noda M, et al. 1989. Structural parts involved in activation and inactivation of the sodium channel. Nature 339:597-603 Suzuki DT, Grigliatti T, Williamson R. 1971. Temperature-sensitive mutations in Drosophila melanogaster. VII. A mutation (para-ts) causing reversible adult paralysis. Proc Natl Acad Sci U S A 68:890-3 Tan J, Liu Z, Nomura Y, Goldin AL, Dong K. 2002a. Alternative splicing of an insect sodium channel gene generates pharmacologically distinct sodium channels. J Neurosci 22:5300-9 Tan J, Liu Z, Tsai TD, Valles SM, Goldin AL, Dong K. 2002b. Novel sodium channel gene mutations in Blattella germanica reduce the sensitivity of expressed channels to deltamethrin. Insect Biochem Mol Biol 32:445-54 Terlau H, Heinemann SH, Stuhmer W, Pusch M, Conti F, et al. 1991. Mapping the site of block by tetrodotoxin and saxitoxin of sodium channel ll. FEBS Lett 293:93-6 Thackeray JR, Ganetzky B. 1994. Developmentally regulated alternative splicing generates a complex array of Drosophila para sodium channel isoforms. J Neurosci 14:2569-78 Thackeray JR, Ganetzky B. 1995. Conserved alternative splicing patterns and splicing signals in the Drosophila sodium channel gene para. Genetics 141:203-14 55 Ulbricht W. 2005. Sodium channel inactivation: molecular determinants and modulation. Physiol Rev 85:1271-301 Vais H, Williamson MS, Devonshire AL, Usherwood PN. 2001. The molecular interactions of pyrethroid insecticides with insect and mammalian sodium channels. Pest Manag Sci 572877-88 Vassilev P, Scheuer T, Catterall WA. 1989. Inhibition of inactivation of single sodium channels by a site—directed antibody. Proc Natl Acad Sci U S A 86:8147-51 Vassilev PM, Scheuer T, Catterall WA. 1988. Identification of an intracellular peptide segment involved in sodium channel inactivation. Science 241:1658-61 Wang SY, Wang GK. 2003. Voltage-gated sodium channels as primary targets of diverse lipid-soluble neurotoxins. Cell Signal 15:151-9 Warmke JW, Reenan RA, Wang P, Qian S, Arena JP, et al. 1997. Functional expression of Drosophila para sodium channels. Modulation by the membrane protein TipE and toxin pharmacology. J Gen Physiol 110:119- 33 West JW, Patton DE, Scheuer T, Wang Y, Goldin AL, Catterall WA. 1992. A cluster of hydrophobic amino acid residues required for fast Na(+)—channel inactivation. Proc Natl Acad Sci U S A 89:10910-4 Wu CF, Ganetzky B. 1980. Genetic alteration of nerve membrane excitability in temperature-sensitive paralytic mutants of Drosophila melanogaster. Nature 286:814-6 Yu FH, Westenbroek RE, Silos-Santiago l, McCormick KA, Lawson D, et al. 2003. Sodium channel beta4, a new disulfide—linked auxiliary subunit with similarity to beta2. J Neurosci 23:7577-85 Zhao Y, Yarov-Yarovoy V, Scheuer T, Catterall WA. 2004. A gating hinge in Na+ channels; a molecular switch for electrical signaling. Neuron 41 :859-65 Zhou W, Chung I, Liu Z, Goldin AL, Dong K. 2004. A voltage-gated calcium- selective channel encoded by a sodium channel-like gene. Neuron 42: 1 01 -1 12 56