Abstract
Ion channels are vital contributors to cellular communication in a wide range of organisms, a distinct feature that renders this ubiquitous family of membrane-spanning proteins a prime target for toxins found in animal venom. For many years, the unique properties of these naturally-occurring molecules have enabled researchers to probe the structural and functional features of ion channels and to define their physiological roles in normal and diseased tissues. To illustrate their considerable impact on the ion channel field, this review will highlight fundamental insights into toxin-channel interactions as well as recently developed toxin screening methods and practical applications of engineered toxins.
Keywords: Animal toxin, Voltage-gated ion channel, Transient receptor potential channel, Toxin engineering, Screening approaches
1. Introduction
Voltage-gated ion channels regulate the ion permeability of the cell membrane and as a result, generate electrical signals that disseminate vital information across the human body1. Evolution has endowed venomous animals and poisonous plants with the ability to exploit this excitatory role by producing toxins that modify ion channel opening or closing (i.e. gating) with the goal of incapacitating prey or defending against predators2. Historically, toxins from scorpion, spider, sea anemone, cone snail, snake, frog, puffer fish, and insect venoms have been used to gain insights into the function, structure, and pharmacological sensitivities of various members of the voltage-gated ion channel family3 including potassium (Kv), sodium (Nav), and calcium (Cav) channels which constitute the main topic of this review. In addition, recent structural advances in the Transient Receptor Potential (TRP) channel field were made possible, in part, by the availability of a unique peptide isolated from tarantula venom that traps the channel in a distinct conformation4; 5; 6. Animal toxins have also contributed to the generation of essential insights into membrane proteins other than voltage-gated ion channels such as acid-sensing7; 8, mechanosensitive9, and chloride ion channels10; acetylcholine11, NMDA12, and G-protein coupled receptors13; and Na+/K+ ATPase14.
In general, toxins that interfere with voltage-gated ion channel function do so through two mechanisms: pore-blocking toxins inhibit ion flow by binding to the outer vestibule or within the ion conduction pore15; 16 whereas gating-modifier toxins interact with a channel region that alters conformation during opening or inactivation to influence the gating mechanism17; 18; 19. As such, gating-modifier toxins constitute powerful tools for researchers seeking to address the unique challenges associated with voltage-gated ion channel voltage sensors as they undergo complex conformational changes during channel activation and inactivation. As illustrated in the next sections, knowledge on the precise working mechanism of toxins is crucial to help elucidate ion channel function. Since many reviews have already summarized a large body of toxin work, this review will illustrate the considerable impact of toxins on the ion channel field by highlighting pioneering experiments that resulted in fundamental insights into toxin-channel interactions as well as potential applications of toxins or toxin-derived compounds. All toxins mentioned in this review are summarized in Table 1.
Table 1. Overview of toxins discussed in this review.
Shown is the organism in which the toxin is found as well as its primary molecular target(s) and binding site(s). Putative binding sites are indicated in parentheses. Note that although tetrodotoxin is found in numerous venomous animals (e.g., pufferfish and blue-ringed octopus); it is actually produced by symbiotic bacteria (largely Pseudoalteromonas, Pseudomonas and Vibrio species). Spider names are from http://research.amnh.org/iz/spiders/catalog/.
| Animal | Species | Toxin | Target | (Putative) binding site |
|---|---|---|---|---|
| Cone snail | Conus geographus | ω-conotoxin GVIA | Cav channels | Pore |
| Conus magus | ω-conotoxin MVIIA | Cav channels | Pore | |
| Conus magus | α-conotoxin MII | nAChR | α3/α6 subunit | |
| Dinoflagellate | Various dinoflagellates/cyanobacteria | Saxitoxin (STX) | Nav/Kv channels1 | Pore |
| Fish/octopus | Produced by symbiotic bacteria | Tetrodotoxin (TTX) | Nav channels | Pore |
| Scorpion | Leiurus quinquestriatus | Charybdotoxin (CTX) | Kv channels | Pore |
| Leiurus quinquestriatus | Agitoxin2 | Kv channels | Pore | |
| Leiurus quinquestriatus | LqTX | Nav channels | Voltage sensor | |
| Androctonus australis Hector | AaHII | Nav channels | Voltage sensor | |
| Tityus serrulatus | Ts3 | Nav channels | Voltage sensor | |
| Centruroides suffusus suffusus | CssII | Nav channels | Voltage sensor | |
| Parabuthus transvaalicus | Kurtoxin | Cav/Nav channels | Voltage sensor | |
| Sea anemone | Anemonia sulcata | BDS-I | Nav/Kv channels | (Voltage sensor) |
| Stichodactyla helianthus | ShK | Kv channels | Pore | |
| Spider | Grammostola rosea | Hanatoxin (1/2) | Kv/Nav/Cav channels | Voltage sensor |
| Grammostola rosea | ω-grammotoxin SIA | Cav/Kv channels | Voltage sensor | |
| Thrixopelma pruriens | ProTx-I | Nav/Kv/Cav channels/TRPA1 | Voltage sensor | |
| Thrixopelma pruriens | ProTx-II | Nav/Cav channels | Voltage sensor | |
| Stromatopelma calceatum griseipes | SGTx1 | Kv/Nav channels | Voltage sensor | |
| Chilobrachys guangxiensis | JZTX-I | Nav/Kv channels | Voltage sensor | |
| Macrothele gigas spider | Magi5 | Nav channels | (Voltage sensor) | |
| Haplopelma huwenum | HWTX-IV | Nav channels | Voltage sensor | |
| Agelenopsis aperta | ω-agatoxin IVA | Cav channels | Voltage sensor | |
| Hysterocrates gigas | SNX482 | Cav channels | Voltage sensor | |
| Psalmopoeus cambridgei | Vanillotoxins | TRPV1 | (Voltage sensor) | |
| Haplopelma huwenum | DkTx | TRPV1 | Pore | |
| Grammostola rosea | GsMTx-4 | TRPA1/Mechanosensitive channels | (Pore) |
See Ref. 210.
2. Voltage-gated potassium channel toxins
Most voltage-gated potassium (Kv) channels are homotetrameric in nature, with each subunit containing six transmembrane helices (S1-S6): the S1-S4 helices form the voltage-sensing domain whereas the S5-S6 helices of four subunits come together in a circular arrangement to form the potassium ion-selective pore20; 21; 22; 23; 24. Toxins that target Kv channels can do so by interacting with the pore region or particular regions within the voltage sensors25. Pore-blocking toxins have greatly facilitated Kv channel research by enabling purification of novel channels and by providing insights into channel subunit stoichiometry as well as the shape of the extracellular pore region26; 27; 28; 29; 30; 31; 32. A particularly well-studied example is charybdotoxin (CTX), a 37-residue peptide isolated from the venom of the deathstalker scorpion Leiurus quinquestriatus (Fig. 1a)33. CTX exhibits a simple, bimolecular binding mechanism, in which a single toxin molecule inhibits the channel by physically plugging the pore (Fig. 1a)34. Early observations led to the hypothesis that CTX approximates a “tethered potassium ion” by bringing a positive charge close to a potassium ion-binding site near the extracellular side within the pore35. This hypothesis was later proven correct when a lysine was identified as the most important residue for CTX function36. This residue is conserved in all members of the CTX-like toxin family (e.g. agitoxin2) that bind with a similar orientation on the Kv channel and inhibit ion flux through a common mechanism37; 38. Recently, the crystal structure of CTX bound to a Kv channel was elucidated (Fig. 3a), a remarkable achievement that required many hurdles to be overcome39. Similar to what was observed with the solid-state NMR structure of the KcsAKv1.3-kaliotoxin complex40, the structure of the CTX-Kv channel complex revealed that the 27th residue of the toxin, a lysine, indeed makes its way into the pore and ends up close to the outermost of the four binding sites for potassium ions that are responsible for the ion selectivity of the channel. This observation confirmed a previously postulated hypothesis as to how intracellular potassium ions can permeate along the pore and influence the dissociation of toxin bound to the external end of the pore35. The structure also explains why mutant toxins without a lysine at this position are less effective at blocking Kv channels.
Figure 1. Kv channel toxins.
a, Left: ion channel cartoon showing one voltage-sensing domain consisting of S1-S4 segments connected to the S5-S6 segments which make up the pore of the channel. CTX binds to the pore of a Kv channel (indicated in red). Middle: Representative NMR structure of CTX (PDB code 2CRD) including both a transparent surface representation as well as the protein backbone with blue = Lys, Arg, His; red = Glu, Asp; purple = Ser, Thr, Tyr; green = Ala, Ile, Leu, Val, Phe, Trp, Met; and yellow = Cys. The same color coding is used in Figure 2. The toxin is oriented with the functionally important Lys27 residue pointing down. Right: Effect of CTX on the conductance (G)-voltage (V) relationship of Kv2.1 Δ746 (control: black circles) expressed in X. laevis oocytes showing that the toxin (grey circles) completely blocks potassium currents over a wide voltage range. Error bars represent SEM. b, Left: hanatoxin binds to the S3b-S4 paddle motif within Kv channel voltage sensors (indicated in red). Middle: Representative NMR structure of hanatoxin (PDB code 1D1H_A). Right: Effect of hanatoxin on the G-V relationship of Kv2.1 Δ7 expressed in X. laevis oocytes demonstrating that the toxin inhibits potassium currents at relatively mild depolarizations whereas the channel can activate with hanatoxin bound at more depolarized voltages43. Hanatoxin is promiscuous and also inhibits Nav channel activation as shown in Fig. 2. Structure images were created using DSViewer Pro. c, Bar graph plot of normalized hanatoxin apparent affinity (Kd) values for alanine mutations spanning from S1 to S4 in Kv2.147. Results revealed the molecular determinants of the hanatoxin receptor site within the Kv2.1 S3b-S4 paddle motif. The solid line superimposed on the bar graph is a 17-residue window analysis of the Kyte-Doolittle hydrophobicity index.
Figure 3. Structures of animal toxins bound to Kv and TRPV1 channels.
a, X-ray crystal structure of CTX bound to Kv1.2 (PDB code 4JTC), viewed from the extracellular side of the membrane39. Each of the four Kv1.2 subunits is shown in a different color (green, pink, blue, and light orange). A single molecule of CTX (red) is bound to the extracellular surface of the channel. The sidechain of the critical Lys27 residue (red tube) protrudes into the channel pore to form hydrogen bonds with backbone carbonyl oxygens at the top of the selectivity filter, thereby blocking ion conduction. b, Side-on view of the structure of DkTx bound to TRPV1 determined using single particle electron cryomicroscopy (PDB code 3J5Q)6. The four TRPV1 subunits are shown in different colors (green, yellow, pink, and blue), and four bound molecules of resiniferatoxin, a TRPV1 agonist, are shown in red. Two DkTx molecules (pink surface representations) bind at the top of the extracellular face of the channel, with each ICK domain (two per DkTx molecule) nestled at a subunit interface, thereby locking the channel in an activated state.
Unlike pore blockers, gating-modifier toxins interact with the voltage-sensing domain to influence Kv channel opening. An early clue that toxins can inhibit Kv channels through a mechanism other than pore occlusion came from experiments with hanatoxin (Fig. 1b), a 35-residue peptide isolated from the venom of the Chilean rose-hair tarantula Grammostola spatulata, where transfer of the outer vestibule (S5–S6 linker) of Kv2.1 into a toxin-insensitive channel failed to confer toxin sensitivity41; 42. Moreover, when co-applied with a pore blocker such as agitoxin2, hanatoxin displayed the ability to bind concurrently, thereby supporting the notion of a binding site outside of the pore region. Ensuing evidence demonstrating that this class of toxins indeed modifies Kv channel gating by influencing their voltage sensors came from three observations. First, toxin-bound channels still open and conduct ions, but the energy (or voltage-step magnitude) required to open toxin-bound channels is substantially increased (Fig. 1b)41; 42; 43. Second, these toxins have distinct effects on voltage sensor movements, as was observed in gating current measurements44; 45. Third, classic as well as more recent mutagenesis experiments suggest that these toxins interact with defined regions within the voltage sensors (Fig. 1c)20; 46; 47; 48. In particular, gating-modifier toxins helped identify the S3b-S4 helix-turn-helix motif, or voltage-sensor “paddle”, which moves at the protein-lipid interface to drive activation of the voltage sensors and opening of the pore47. Strikingly, this motif can be transplanted into different voltage-gated ion channel isoforms without losing their capacity to interact with toxins20. One intriguing aspect of gating-modifier toxins is their ability to interact with the paddle motif in the resting conformation in which case the voltage sensor is buried within the lipid membrane. However, the amphipathic character observed in many toxin structures (Fig. 1b) is consistent with the notion that membrane partitioning may be required for the toxin to reach the channel49; 50; 51; 52; 53. It will be interesting to determine the docking site of gating-modifier toxins with the paddle motif using mutant cycle approaches or by obtaining high-resolution structures of a gating-modifier toxin bound to a Kv or Nav channel. Still, toxin binding affinities and large-scale production are two important limitations to overcome.
A unique feature of voltage-sensor targeting toxins is that they can interact with different families of voltage-activated ion channels. Such promiscuous behavior has been observed for toxins such as hanatoxin (see Figs. 1 and 2), ω-grammotoxin SIA, ProTx-I, ProTx-II, BDS-I, and SGTx142; 46; 54; 55; 56; 57. The widespread targeting of paddle motifs by animal toxins highlights the pharmacological importance of this part of the voltage sensor.
Figure 2. Nav channel toxins.
Effects of 10nM TTX, 100nM AaHII (an α-scorpion toxin from Androctonus australis Hector)207, 1μM CssII (a β-scorpion toxin from Centruroides suffusus suffusus)208, and 100nM hanatoxin on rNav1.2a channels expressed in X. laevis oocytes. Hanatoxin is promiscuous and also inhibits Kv channel activation as shown in Fig. 1. Left column shows the location within the Nav channel where these toxins bind (red) which is either the pore region for TTX (a) or the voltage-sensing domain for AaHII, CssII, and hanatoxin (b). Middle column displays representative toxin structures (PDB codes for AaHII and CssII are 1PTX and 2LJM, respectively) including both a transparent surface representation as well as the protein backbone with the same residue color coding as in Figure 1. Right column shows rNav1.2a sodium currents elicited by a depolarization to a suitable membrane voltage before (black) and after toxin addition (red). TTX potently blocks currents (a) whereas depending on which voltage sensors are targeted and how they couple to Nav channel gating, scorpion and spider toxins (b) can inhibit channel fast inactivation (domain IV: AaHII), facilitate channel opening by shifting the activation voltage to more hyperpolarized membrane potentials (domain II: CssII), or inhibit channel opening in response to membrane depolarization (domain I/II/IV: hanatoxin).
3. Voltage-gated sodium channel toxins
In contrast to Kv channels, the channel-forming component of the voltage-gated sodium (Nav) channel complex consists of four domains (DI-IV), each with six transmembrane segments (S1-S6), connected by intracellular linkers58. These similar, but non-identical domains each consist of a voltage-sensing region (S1-S4), while the S5-S6 helices from each domain come together to form a sodium ion-selective pore in the membrane. The pore can open when all four voltage sensors move in response to changes in membrane voltage. In general, all four sensors activate following membrane depolarization; however, those in domains I-III are most important for channel opening, whereas the one in domain IV plays a distinctive role in fast inactivation54; 59; 60; 61; 62; 63; 64. Recently, the first crystal structures of bacterial Nav channels were solved65; 66; 67; 68. Although seminal observations were reported, prokaryotic Nav channels are homotetramers whereas their mammalian counterparts are multi-domain monomers. Moreover, key characteristics of mammalian Nav channels such as the amino acid composition of the selectivity filter, fast inactivation gate, and presence of auxiliary subunits differ substantially from prokaryotic variants65; 66; 67; 68.
A few years after Hodgkin and Huxley discovered the fundamental role of Nav channels in electrical signaling69; 70, one of the historically most important ion channel toxins was isolated from the Japanese puffer-fish71. This naturally-occurring marine toxin, tetrodotoxin (TTX), interacts strongly with the Nav channel pore region to occlude the sodium ion permeation pathway (Fig. 2a)72; 73. TTX has been widely used to study the structural and functional properties of Nav channels. For example, after the TTX structure was solved in 196474, Hille exploited this information to predict the diameter of the Nav channel pore, thereby providing unique insights into the molecular structure of this ion channel family15. Later, the low-nanomolar affinity of [3H]-TTX contributed to the isolation of the Nav channel pore-forming subunit from the electric eel75; 76. Likewise, [3H]-saxitoxin (STX) isolated from marine dinoflagellates played a vital role in the purification of a rat brain and skeletal muscle Nav channel isoform, thereby also demonstrating the existence of auxiliary β-subunits77; 78; 79. Moreover, recently synthesized STX variants show promise as a tool to elucidate the role of Nav channels in signal conduction and their dysregulation in specific disease states80; 81. Nowadays, TTX-sensitivity is used to classify the nine mammalian Nav channel isoforms into two groups: TTX-sensitive channels (Nav1.1-Nav1.4, Nav1.6-Nav1.7) are blocked by nanomolar concentrations of TTX whereas Nav1.8 and Nav1.9 are inhibited by millimolar concentrations1. Although inhibition of Nav1.5 requires micromolar concentrations of TTX, the response can be substantially increased by substituting a cysteine in the S5-S6 loop of domain I with tryptophan or phenylalanine82; 83; 84. These aromatic residues support high affinity blockade through the formation of a cation-π interaction with TTX83; 85. The cationic groups also interact with anionic amino acids within the pore of the channel to prevent Na+ flux. Experiments designed to pinpoint residues that reduced the susceptibility of Nav channels to TTX led to the discovery of the DEKA motif which consists of four residues that make up the Nav channel selectivity filter: aspartate (DI S5-S6 loop), glutamate (DII S5-S6 loop), lysine (DIII S5-S6 loop), and alanine (DIV S5-S6 loop)86; 87.
Similar to Kv channels, animal venoms also contain toxins that target Nav channel voltage sensors to disrupt channel function. Well-studied examples include members of the long-chain scorpion toxin family which, in general, interact with the extracellular loops between S3 and S4 to stabilize the Nav channel voltage sensors in the activated or resting state17; 18; 19; 88; 89. Based on their functional effects, two classes of Nav channel scorpion toxins have been defined90. First, early antibody and photo-affinity labeling studies indicated that the α-scorpion toxin LqTX interacts with the S5-S6 loops of domains I and IV in rat neuronal Nav channels91; 92. Later, mutagenesis experiments revealed a primary interaction with the S3-S4 paddle motif within domain IV, as well as a secondary binding site within the domain I S5-S6 loop and domain IV S1-S2 loop93; 94. Because of their binding locus, α-scorpion toxins as well as the functionally similar sea anemone toxins have been used extensively to elucidate the role of the domain IV voltage sensor in Nav channel gating (Fig. 2b). For example, a recent study using Ts3 from the Brazilian yellow scorpion Tityus serrulatus on native Nav channels within GH3 cells in concert with fluorescently labeled voltage-sensing domains, demonstrated that the toxin's inhibitory effect on movement of the domain IV voltage-sensing domain results in inhibition of fast inactivation and facilitation of recovery from fast inactivation95.
A second class of Nav channel gating-modifier toxins found in scorpion venom are the long-chain β-scorpion toxins which induce sub-threshold channel opening by shifting the threshold for channel activation to more negative membrane potentials (Fig. 2b)54; 96; 97. Distinct from α-scorpion and sea anemone toxins, β-scorpion toxins primarily target the paddle motif within the domain II voltage sensor and stabilize it in an activated state98; 99. As a result, subsequent depolarizations require the transition of fewer voltage-sensing domains, resulting in the hyperpolarized voltage-dependent activation observed in toxin-bound channels.
In contrast with scorpion and sea anemone toxins, exploration of the mechanism through which spider toxins belonging to the inhibitor cysteine knot (ICK) family interact with mammalian Nav channels is a more recent phenomenon. Depending on which voltage sensors are targeted and how they couple to the Nav channel gating process, spider toxins can have three diverse effects on ion channel function100. Most commonly, these toxins inhibit channel opening in response to membrane depolarization (e.g. hanatoxin (Fig. 2b), ProTx-I, ProTx-II)42; 55. A second possibility for these toxins is to prevent fast inactivation by impairing movement of the domain IV voltage sensor (e.g. SGTx1, JZTX-I)54; 101. Finally, and similar to β-scorpion toxins, spider ICK toxins can facilitate channel opening by shifting the activation voltage to more hyperpolarized membrane potentials (e.g. Magi5)102.
Recently, the identification of an S3b–S4 paddle motif within each of the four Nav channel voltage sensors significantly advanced our insights into the multifaceted working mechanism of spider toxins54. In this study, it was demonstrated that each of the four paddle motifs can interact with toxins from spiders, scorpions, and sea anemones, and that multiple paddle motifs can be targeted by a single toxin. These novel insights also led to the identification of the first tarantula toxin (ProTx-I) active on Nav1.9, an enigmatic Nav channel isoform predominantly expressed in nociceptive dorsal root ganglion neurons103. The taxonomically related tarantula toxin ProTx-II targets Nav1.7, a Nav channel isoform implicated in various debilitating pain syndromes, with an affinity that is >100-fold higher than other Nav channel isoforms104. As is the case with HWTX-IV from the Chinese bird spider Selenocosmia huwena, ProTx-II preferentially targets the domain II voltage sensor in Nav1.7105; 106; 107. However, ProTx-II also binds to the domain IV paddle motif in this channel subtype, thereby resulting in a slowing of fast inactivation. Overall, toxins isolated from spider venom have already proven to be valuable tools for probing the structure and functional mechanisms of Nav channels. Moreover, future opportunities may arise for the use of these small and stable peptides as therapeutic drugs.
4. Voltage-gated calcium channel toxins
The channel-forming subunit (α1) of voltage-activated calcium channels (Cav) shares a similar architecture with the corresponding part in Nav channels: four homologous domains (DI-DIV), each containing a voltage-sensor (S1-S4 segments) and two transmembrane segments (S5-S6) that contribute to the pore structure108; 109. Like Nav channels, α1 can function as a standalone subunit that determines many of the biophysical and pharmacological properties of the channel. However, Cav channels are normally formed by association of α1 with three auxiliary subunits (β, α2δ, and γ) that regulate channel function and expression108; 110. In contrast to the other voltage-gated ion channels, Cav channels have the distinctive property of allowing voltage-controlled membrane passage of calcium ions, a key intracellular signaling factor. As the permeant ion has a positive reversal potential, the opening of Cav channels results in further membrane depolarization, which allows them to function as an action potential generating mechanism, and to control the firing activity of excitable cells108. Thus, Cav channels are involved in physiological processes where they couple calcium signaling with membrane excitability, such as hormone and neurotransmitter release, and initiation of electrical activity-dependent transcriptional events111. Historically, Cav channels were classified based on how calcium currents activated in response to membrane depolarization. Two large families were found: low-voltage (LVA: T-type) and high-voltage-activated (HVA: L-, N-, P/Q-, and R-types) channels112. Subsequent biochemical and pharmacological studies and molecular cloning identified the molecular identities that form Cav channels. Thus, the L-type corresponds to the Cav1 subfamily113, the N-, P/Q-, and R-types correspond to Cav2112, and the T-type corresponds to Cav3114. Extensive alternative splicing of the ten known α1 genes, together with combinations of several β-subunit types and posttranslational modifications, further enhance the molecular and functional diversity of Cav channels115.
Organic molecules and especially toxins have been instrumental in discovering the many members of the Cav family as well as in understanding their molecular structure and function. For example, the L-type channel was isolated and characterized on the basis of its sensitivity to dihydropyridine116; 117; 118. However, most subsequent discoveries were enabled by the identification of two groups of toxins: the ω-conotoxin family of pore blockers and the functionally heterogeneous ω-agatoxin family of pore blockers and gating modifiers. The first ω-conotoxin, GVIA, was discovered in the venom of the cone snail Conus geographus and is still one of the most widely used pharmacological agents in the study of Cav channels119; 120; 121. GVIA is selective for N-type (Cav2.2) channels and was initially employed to inhibit neurotransmitter release, thus functionally linking N-type channels to synaptic transmission120; 122. Later, GVIA was used to determine the molecular identity of the human N-type channel, by showing that a heterologously expressed channel consisting of α1b, β2, and α2δ clones was able to generate a current with kinetic properties similar to the N-type current endogenously expressed in neurons123; 124. Evidence supporting a pore-blocking mechanism for GIVA comes from several observations119; 121; 125; 126. First, early structure-function studies relied on GVIA125; 127 to determine the architecture of the outer vestibule of the N-type channel. Consequently, chimeric constructs between the GVIA-sensitive (N-type) and GVIA-insensitive (P/Q-type, Cav2.1) channels showed that the toxin binds to the S5–S6 linker region of domain III of the α1 subunit125 suggesting that GVIA likely acts via physical occlusion of the pore. Second, increasing the external concentration of the permeant ion greatly decreased the efficiency of toxin block, indicating that the toxin competes with divalent cations bound at the locus of selectivity128; 129. Third, residues in close proximity to the region identified by Ellinor et al.125 also contribute to block of N-type Cav channels by GVIA since a single amino acid substitution in this region is sufficient to alter the reversibility of toxin block126. Fourth, GVIA block is prevented by ω-agatoxin IIIA binding, which partially occludes the pore to reduce channel conductance130. Of note is that the toxin binds to the channel with very high affinity; unless the channel has been inactivated, in which case the interaction is much weaker127. The ability of GVIA to discriminate between activated and inactivated states of the channel suggests that the process of inactivation is accompanied by a conformational change in the outer vestibule127. Evidence that GVIA may also alter the gating properties of N-type channels stems from work by Jones et al.131 in which they report that the toxin affects inactivation-induced immobilization of N-type channel gating currents at depolarized voltages. This hypothesis was substantiated by Yarotskyy and colleagues who showed that GVIA induced a 10mV depolarizing shift in the gating charge versus voltage relationship as well as a decreased off-gating current time constant and a smaller on-gating current132.
The most relevant member of the ω-agatoxin family133 is IVA, a gating modifier toxin isolated from the funnel-web spider Agelenopsis aperta. IVA is specific for P/Q-type channels and it has been used to isolate P-currents in Purkinje neurons134; 135 and Q-currents (also with ω-conotoxin MVIIC136) in granular cerebellar neurons137; 138; 139. The subunit composition of P/Q channels was determined by immunoprecipitation of the channel complex labeled with IVA, which suggested that α1, α2δ, and β subunits but not γ subunits form the channel140; 141; 142. IVA has 10 times higher affinity for P-channels, and thus it can distinguish between P- and Q-subtypes in mammalian tissues. It is known now that the P- and Q-subtypes are the result of alternative splicing in the S3-S4 linker of domain IV143, which has also been identified as a toxin binding site144; 145; 146. However, recent evidence suggests that this alternative splicing cannot fully explain the difference in affinity for IVA147. Other factors, such as expression system-dependent toxin responses, the identity of the auxiliary subunit(s), and post-translational channel modifications may also be involved in generating the complete P- and Q-channel phenotypes143; 148.
The HVA channels are targeted by multiple members of the ω-conotoxin and ω-agatoxin families and by related toxins, such as ω-grammotoxin SIA149 and SNX482150. In contrast, there are fewer known toxins that target LVA channels. The first discovered toxin that acts with high-affinity against T-type channels is kurtoxin, isolated from the venom of the scorpion Parabuthus transvaalicus. Kurtoxin is a gating modifier that affects the kinetics of activation and inactivation of the channel. It has been proposed that the toxin interacts with the domain IV voltage-sensor of the T-type channel151; 152. Recent studies identified other toxins (e.g., ProTx-I and ProTx-II) that inhibit T-type channels153; 154. However, their affinity and effects on channel gating are different compared with those of kurtoxin. For example, ProTx-II modifies both the activation and deactivation kinetics but has no effect on inactivation153. Determining the binding sites of these toxins will be useful to dissect out the role of individual voltage sensors in the gating mechanism and the structural aspects of inactivation in T-type channels.
5. Transient Receptor Potential channel toxins
Transient Receptor Potential (TRP) ion channels are tetrameric non-selective cation channels that are architecturally similar to the members of the voltage-gated ion channel family155; 156. Although some TRP channels display modest voltage-activation157; 158, they are functionally quite distinct from voltage-gated ion channels. One defining characteristic of TRP channels is that they are activated by a diverse range of stimuli including temperature, small organic compounds, and mechanical stress159; 160. The mechanistic underpinnings of this polymodal gating are poorly understood and the lack of selective pharmacological tools has greatly hampered mechanistic studies on TRP channels. Therefore, recent reports on the discovery of animal toxins that selectively modulate the activity of TRP channel subtypes have given a fresh lease of life to efforts focused on understanding the mechanism of action of these enigmatic ion channels. Additionally, TRP channels are believed to play important roles in pain signalling161; 162 giving rise to the intriguing possibility that selective toxin modulators of these channels may have therapeutic potential. Indeed, reports on modulation of TRPV1 and TRPA1 channels by animal toxins have energized the field tremendously and the major findings of these studies are summarized below.
Also known as the vanilloid receptor, TRPV1 is the most extensively studied TRP channel163; 164. It is activated by a diverse range of small molecules including capsaicin, the active ingredient of chilli peppers that is responsible for their “hot” sensation, and also by temperatures >42°C. The first report on the activation of TRPV1 by animal toxins was published in 2006 by Julius and co-workers who reported the isolation of three toxins from the venom of the Trinidad chevron tarantula Psalmopoeus cambridgei and demonstrated that these toxins activated rat TRPV1165. Furthermore, these “vanillotoxins” had no effect on the activity of other tested TRP channels. The most potent of these toxins, VaTx3, activated TRPV1 with an EC50 of 0.32 μM. The least potent TRPV1 activator among these three vanillotoxins, VaTx1, was found to also inhibit Kv2.1. This inhibition was characterized by a rightward shift in the conductance-voltage relationship of Kv2.1, an effect that mirrors that of voltage sensor-targeting toxins such as hanatoxin that bind to voltage-gated channel paddle motifs25; 42.
In 2010, another landmark study reported the discovery of a so-called double-knot toxin (DkTx) which was isolated from the venom of the Chinese bird spider, Haplopelma huwena and demonstrated to be a potent TRPV1 agonist4. This toxin consists of two ICK domains attached via a linker. Each ICK domain has the ability to individually activate TRPV1, albeit with much lower potency compared to DkTx4; 166. Moreover, channel activation by the separated domains is readily reversible whereas unbinding of DkTx is extremely slow, resulting in almost irreversible activation of the channel. The concentration-response curve for single unit-activation of TRPV1 yields a Hill coefficient of >1 suggesting that several DkTx molecules bind to the channel in a cooperative fashion. A high-yielding E. coli expression system for the heterologous production of DkTx has been developed and has provided access to milligram quantities of the toxin, thereby facilitating structural and functional studies4; 5; 166. Recent single-particle electron cryo-microscopy studies provided a 3.8 Å resolution structure of the DkTx-TRPV1 channel complex which also has a small molecule agonist, resiniferatoxin, bound to the channel (Fig. 3b)6. This structure revealed that two molecules of DkTx bind to one molecule of TRPV1 at the subunit interface in the pore domain (the S5-S6 region). Comparison of this structure with that of the channel without bound ligand167 provides valuable insights into the conformational changes that are involved in channel opening upon toxin binding. It appears that the S1-S4 domain of TRPV1 remains relatively static during gating whereas the mobile regions in TRPV1 are the pore helix and the outer loop connecting the S5 and S6 helices. These observations paint a fascinating picture of the conformational changes involved in TRPV1 gating which appear to be in sharp contrast to those involved in gating of voltage-activated ion channels where the S1-S4 domain is believed to undergo large conformational changes upon application of voltage stimuli and the S5-S6 pore domain remains relatively static24; 168. At this stage, however, these ideas are mere hypotheses that will no doubt be tested thoroughly in the near future by detailed biophysical studies.
The TRPA1 channel is characterized by the presence of seventeen repeats of the ankyrin protein motif169; 170. It is activated by plant products such as icilin, allicin, and isothiocyanates and is believed to be activated by noxious cold and mechanical stress171; 172; 173; 174; 175. In 2007, Schaefer and co-workers reported that an ICK toxin isolated from G. spatulata, GsMTx-4, activates human TRPA1 expressed heterologously in HEK293 cells176. This toxin is also a blocker of mechanosensitive ion channels9. Recently, ProTx-I, a promiscuous inhibitor of voltage-gated ion channels54, was shown to be an antagonist of TRPA1177. In this work, the authors employed the recently developed tethered-toxin approach178; 179 to screen ~100 peptide toxins as potential modulators of TRPA1, which led to the discovery of ProTx-I as a TRPA1 inhibitor.
The discovery of animal toxins that selectively modulate TRP channel function has tremendous implications in TRP channel biology. They are likely to be powerful mechanistic tools for understanding the molecular details of TRP channel gating, in much the same way that both pore-blocking and paddle-targeting ICK toxins have contributed to our understanding of gating in voltage-activated channels. Moreover, the use of these toxins as therapeutic molecules targeting TRP channels remains an unexplored area of research and may be worth pursuing given the success of a cone snail toxin (ω-conotoxin MVIIA) targeting spinal populations of Cav2.2 that obtained FDA-clearance for treatment of intractable chronic pain180.
6. High-throughput screening for novel animal toxins targeting ion channels
Most ion channel toxins reported to date, including those described in this review, were discovered by serendipity or low-throughput characterization of individual venom components. Conversely, most venom peptides are now discovered through holistic analysis of venoms using a combination of advanced proteomic methods and sequencing of venom-gland transcriptomes181. The ability to produce these peptides via recombinant expression182 or chemical synthesis183 means that a much larger number of animal toxins are now available for drug discovery efforts and, moreover, it provides access to toxins from miniature venomous animals that would otherwise be inaccessible using conventional bioassay-guided venom fractionation. However, the growing number of available toxins combined with an increasing interest in venom-based ion channel drug discovery has created a need for higher-throughput discovery pipelines181.
The basic requirements of high-throughput screening (HTS) include high-assay sensitivity and accuracy as well as high-assay robustness and reproducibility. This is particularly important with respect to the screening of crude venoms or partially purified venom fractions. In contrast to combinatorial chemical libraries, which often comprise many thousands if not millions of compounds, natural product libraries frequently consist of crude or partially purified mixtures of compounds with diverse biological effects. This reduces the resource requirements for screening, but at the same time introduces the potential for interference from non-target-specific effects from other components of the venom. Thus, while the traditional goal for HTS in the pharmaceutical industry has been to increase screening capacity through automation and miniaturization of assays, HTS in the context of venom-based drug discovery arguably requires greater emphasis on data quality181.
Fluorescence-based assays are commonly used for high-throughput ion channel screens184; 185. In these assays, the transmembrane flux of ions resulting from channel activation is detected by changes in fluorescence using either ion-specific or membrane-potential-sensitive dyes. The change in fluorescence is detected using specialized plate readers (e.g., FLIPRtetra). Calcium-sensitive dyes186 have been used most frequently in high-throughput ion channel screens184. A significant limitation with these and other non-electrophysiology-based HTSs is the lack of voltage control, which is important when screening for toxins that interact with voltage-gated ion channels. As discussed above, many ion channel toxins are gating modifiers that have highest affinity for specific states of the channel. Without voltage control, the conductance state of the channel cannot be controlled, making it difficult to detect toxins with certain modes of action, such as those that cause use-dependent inhibition. Use-dependent toxins are of particular interest for modulating channels where subtype-selective binding is difficult to achieve but mechanistic selectivity yields an acceptable therapeutic window (e.g., use-dependent Nav channel inhibitors that target rapidly firing sensory neurons might be useful analgesics).
The ‘gold standard’ for determination of the mechanism of action of ion channel modulators is electrophysiology, and recent advances include the development of automated platforms for single-cell electrophysiology studies, thus increasing throughput184. Several planar array-based automated electrophysiology systems have been developed to incorporate the precision and accuracy of manual patch-clamp experiments184. Of these systems, the PatchXpress® (Axon Instruments), QPatch™ (Sophion) and Patchliner® (Nanion) platforms have gained acceptance as ion channel drug-discovery platforms in the pharmaceutical industry184. These automated patch-clamp instruments allow screening for state-dependent inhibitors187.
In order to rationally improve the potency and selectivity of ion channel toxins, it is essential to map the epitopes that mediate their interaction with the channel of interest (i.e., the toxin pharmacophore). Structure-activity relationships for peptides are traditionally developed in the first instance via alanine scanning mutagenesis, where a panel of toxin mutants is produced in which each residue is individually replaced with Ala and the effect on toxin function is examined188. Although this approach has been used to map the pharmacophore of several ion channel toxins51; 53; 189; 190; 191, it is laborious and time consuming as a full panel of alanine mutants must be produced by chemical or recombinant methods. A much faster and cheaper alternative is to perform the entire alanine scan using the recently reported tetheredtoxin approach177. In this method, oocytes are injected with mRNA encoding the channel of interest plus mRNA encoding native toxin or one of the alanine mutants. The toxin construct encodes an N-terminal signal sequence to ensure that the toxin is exported from the cell and a C-terminal glycosylphosphatidylinositol (GPI) recognition sequence that directs the oocyte to covalently attach a GPI anchor to the toxin178; 179. Thus, after production and export, the peptide toxin is tethered to the extracellular surface of the oocyte via the C-terminal GPI anchor177. Inclusion of a long hydrophilic linker between the C-terminus of the peptide and the GPI anchor provides sufficient flexibility for the toxin to access the co-expressed channel, while tandem Myc epitopes incorporated into the linker region facilitate quantification of toxin expression177. Because this approach obviates the need for production and purification of toxin mutants, it reduces the time period for performing an alanine scan from months to as little as one week. As for any alanine scanning mutagenesis approach, false positives will be obtained if the tethered toxin is not correctly folded. Thus, it is important to subsequently produce and test the function and folding of the small number of mutants identified as hits in the tethered-toxin screen. Despite this requirement, the greatly improved throughput of this approach facilitates the mapping of pharmacophores across multiple channels if a toxin has polymodal pharmacology, thereby enabling the rational engineering of toxin selectivity. For example, this approach was recently used to map the pharmacophores on ProTx-I that mediate its interaction with TRPA1 and NaV1.2. These pharmacophores were found to be distinct but overlapping, which allowed rational design of analogues that are selective for TRPA1 or Nav1.2177.
7. Toxin engineering: a rationale
The functional properties of toxins found in animal venoms are generally suitable for use as ligands for ion channels, since these have been honed by evolution to act effectively in a native physiological context. Although many toxins can be utilized in their unaltered state, there are a number of reasons for modifying them. To provide an intellectual framework for this section, we discuss the rationale for “toxin engineering”.
One reason for initiating a toxin-engineering project is to alter toxin selectivity. For example, it may be desirable to shift toxin specificity from its preferred molecular target to another one. Alternatively, a native toxin may not act with sufficient selectivity or potency and peptide engineering is required to enhance specificity for the particular channel isoform being investigated. For biomedical applications or for diagnostic and therapeutic purposes, a wide variety of toxin engineering protocols can be envisioned that would alter 1) the binding properties of the peptide, 2) its ability to cross membranes, 3) its stability under physiological conditions, or 4) its pharmacodynamic properties192. Since this is a vast and rapidly moving field193; 194; 195, it falls beyond the scope of this review. Instead, we concentrate on toxin engineering for improving molecular specificity, an important issue for those toxins to be used as tools for investigating basic ion channel function.
Most animal toxins have been optimized by natural selection to efficiently target a single physiologically-relevant protein target present in a specific animal (e.g., the prey, predators, or competitors of the venomous species that evolved the toxin). If selection has indeed occurred to target a specific ion channel, would it really be productive to further modify a toxin? An example of a specific scenario that would require toxin engineering is as follows. Researchers employ specific toxins to investigate circuits present in mammalian systems (and less frequently, other vertebrates, or in non-vertebrate model systems such as Drosophila). However, the actual molecular target of a venom peptide may be an ion channel in specific phylogenetically-distant insects. Nonetheless, this target may be a homolog of an ion channel of interest in a specific mammalian circuit. Because of the evolutionary distance, there is not necessarily a 1:1 correspondence. For example, there may be a single physiologically-relevant molecular target in the insect, yet in mammalian systems that ion channel may have diverged into two closely-related homologous ion channel isoforms. Consequently, this venom peptide may target both mammalian isoforms with equal affinity. To be a suitable research tool, this toxin would need to be engineered to be specific for only one of the two derivative isoforms in the mammalian circuitry. In the sections that follow, we examine two actual cases of toxin engineering that were carried out with the goal of developing powerful tools for studying ion channel function. In both cases, the molecular targeting specificity of a native toxin was altered to be appropriate for mammalian applications.
A first example of toxin engineering concerns ShK196, a sea anemone toxin from Stichodactyla helianthus, that has been used to investigate the molecular mechanisms underlying autoimmunity. Early investigations by Chandy and coworkers established the expression of specific Kv channel isoforms in T-cells. The specific T-cell subclass implicated in the pathology of autoimmune disease requires the expression of KV1.3, a member of the Shaker subfamily of K v channels197. This switch in gene expression is essential for amplification of the T-cell population responsible for diverse types of autoimmune disease. In the search for ligands specific for Kv1.3, ShK was found to block this particular Kv channel isoform with high affinity. However, the native toxin also inhibited the closely related KV1.1 isoform. Subsequently, Chandy and coworkers carried out a toxin-engineering program in which they synthesized 38 analogs to improve ShK selectivity towards KV1.3 (see Chi et al., 2012 for a review198). As a result of these efforts, several analogs were identified that showed much improved selectivity for KV1.3. It is worth mentioning that in this particular case, the toxin engineering strategy did not involve substitution of the normal complement of amino acids. Instead, ShK was chemically derivatized at the N-terminus with an L-phosphotyrosine attached to an aminoethyloxy-acetyl linker and was also amidated at the C-terminus. Next, the Kv1.3 selective analogs were shown to be efficacious in various animal models of autoimmune disease. One of these analogs, ShK-186, is currently under drug development by Kineta Inc.; ShK-186 recently completed Phase Ia clinical trials and it will soon enter Phase 1b trials for treatment of psoriatic arthritis. Preliminary work using a variety of animal models indicates the potential of ShK derivatives for treating diverse types of autoimmune disease such as type I diabetes, rheumatoid arthritis, and multiple sclerosis.
Given that the KV1.3 isoform is expressed in many tissues including neurons, it was surprising to learn that the engineered sea anemone peptide is clinically effective without having a serious side-affect profile. Chandy and coworkers have a potential explanation for this phenomenon: there is a metalloproteinase expressed in neurons, which has a domain that binds to the homomeric KV1.3 channel but not to heteromeric combinations199; 200. The implication is that in neurons, KV1.3 channel homomers cannot be expressed on the plasma membrane; however, immune cells lack the metalloproteinase and as such the T-cells implicated in autoimmunity may be the only significant cell population with the homomeric KV1.3 channel expressed on the membrane surface. As a result, the engineered ShK variant has much more specificity than would have been anticipated purely from KV1.3 gene expression patterns. Overall, this case nicely illustrates the advantages of having a highly-specific ligand for understanding the role of a molecular target in normal physiology and in pathophysiology.
A second example of toxin engineering relates to nicotinic acetylcholine receptors (nAChRs) present in the dopaminergic circuitry (for a review, see Quik and McIntosh, 2006201; Olivera et al., 2008202; Azam and McIntosh, 2009203). Although diverging from the focus of the review on voltage-gated ion channels, we present this example to illustrate how toxin engineering can lead to surprising insights into ion channel function. The biomedical rationale of this study is the observation that heavy smokers appear to be protected from Parkinson's disease, perhaps because the nicotine in tobacco provides a protective effect204. In order to understand the molecular machinery underlying this phenomenon, it is desirable to define which nicotinic receptor isoforms might mediate the protective effect of nicotine.
The initial data that triggered this investigation is shown in Figure 4. A radiolabeled Conus peptide, α-conotoxin MII from the fish-hunting cone snail Conus magus (the Magician's cone), was shown to label the same region of the striatum as a ligand that binds the dopamine transporter. Since α-conotoxins are nAChR antagonists, this observation suggested that the target of the peptide was present in the dopaminergic circuitry. In order to determine the selectivity of α-conotoxin MII, cloned nAChR subunits were expressed and assayed in Xenopus oocytes and ensuing experiments revealed that α-conotoxin MII inhibits nAChRs containing α6 or α3 subunits. This is not entirely unexpected, since these two α-subunits are the most closely related of all mammalian nAChR subunits. However, employing the toxin to identify the precise composition of the nAChRs present in the striatal dopaminergic circuitry would be impossible. An additional complication is that functional nAChRs are pentameric, with most being heteromeric combinations of two α- and three non-α-subunits. Thus, not only was there a problem in differentiating between α3 and α6-containing nAChRs, but also in determining the identity of the other subunits. One approach to remedy this limitation was to engineer an α-conotoxin MII variant with greater subunit selectivity. This was accomplished by synthesizing a series of single-substitution analogs, followed by combining those substitutions that had higher affinity for α6-containing nAChRs compared to those containing α3205. Using an engineered peptide with two amino acid substitutions, it was readily established that the dopaminergic circuitry in the striatum had α6-containing (and not α3) nAChRs. In combination with knockout mice studies193, the receptor was defined as having a (α6)2(β2)2β3 composition.
Figure 4. α-conotoxin MII interacts with nicotinic receptors.
Top panel shows autoradiography carried out using radiolabeled α-conotoxin MII, as well as a radiolabeled ligand specific for the dopamine transporter. There is a striking overlap between labeling with the toxin and with the transporter ligand. The left panel shows (1) the sequences of native α-conotoxin MII that was characterized by Cartier et al.209, from the venom of Conus magus, and (2, 3) two analogs developed through toxin engineering (the H9A, L15A analog and the E11A analog, respectively)205. The potential targets of α-conotoxin MII and its derivatives are shown on the right-hand panel. The native peptide (1) is an antagonist of all three molecular subtypes (A, B, C); the middle analog (2) differentiates between the α3 (A) and α6-containing (B, C) isoforms, and only antagonizes the latter. Analog 3 has the highest affinity for the C isoform, which has both an α6 and an α4 subunit. Thus, toxin engineering allowed identification of the α6-containing isoforms (B, C) in the dopaminergic circuitry.
In the course of these studies, an unexpected observation was made: when radiolabeled α-conotoxin MII was displaced by either the native variant or by the analog that could distinguish between α6 and α3, a single displacement curve was obtained. However, one analog of α-conotoxin MII (shown in Fig. 4) had a biphasic displacement, suggesting that there were two isoforms of the α6-containing nAChR present in the striatum. Follow-up experiments indeed established that the two α6-containing isoforms in the striatum are likely to be those shown in the figure, i.e., one isoform containing (α6)2(β2)2β3 and a second isoform containing α6α4(β2)2β3. These studies also indicated that the high-affinity isoform that contains an α6 subunit disappears early in the progression to a disease state in a primate model of Parkinson's disease206. Thus, engineering native α-conotoxin MII enabled the identification of the nAChR subtype in the dopaminergic circuitry as well as the possible connection of these nicotinic receptors with the progression of the disease state.
4. Conclusion
Animal toxins have been successfully employed for many years to investigate ion channel function. The ability to construct cDNA libraries from animal venom glands together with recombinant/synthetic production methods and high-throughput screening techniques will allow researchers to better exploit animal peptides to investigate the functional aspects of various ion channel isoforms and receptors as well as examine their role in normal and disease states. Moreover, synthetic methods such as cyclization, minimization, and the use of diselenide bridges192; 195 may overcome the challenges that exist between the initial drug discovery stage and the clinical application of these toxins, which will enable their use as therapeutics.
Highlights.
Ion channels are vital contributors to cellular communication and are targeted by toxins found in animal venoms.
For many years, the unique properties of toxins have enabled researchers to probe the structural and functional features of ion channels.
This review highlights fundamental insights into toxin-channel interactions by focusing on representative examples from the voltage-gated sodium, voltage-gated potassium, voltage-gated calcium, and transient receptor potential channel field.
Recent developments in toxin screening methods are discussed.
Practical examples of cases in which toxin engineering was successfully used to address specific ion channel questions are also provided.
Acknowledgements
We would like to thank Kenton J Swartz (NINDS-NIH) for making available the data shown in Figure 1. We would also like to thank Marie-France Martin-Eauclaire and Pierre E Bougis (University of Marseille) for sharing AaHII and CssII for conducting experiments shown in Figure 2 as well as the members of the David Julius's and Yifan Cheng's labs (UCSF) for sharing an image of the TRPV1 structure shown in Figure 3. This publication was supported by the National Institute of Neurological Disorders and Stroke (NINDS) of the National Institutes of Health (NIH) under award number R00NS073797 (FB), by a National Institute of General Medical Sciences Grant GM48677 (BMO), by the American Heart Association (AHA 13SDG14570024) (MM), and the Australian National Health & Medical Research Council (Principal Research Fellowship and Project Grant APP1034958 to GFK).
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
References
- 1.Hille B. Ion channels of excitable membranes. 3 edit 1. 1. Sinauer Associates, Inc.; MA, USA: 2001. [Google Scholar]
- 2.Mebs D. Venomous and Poisonous Animals: A Handbook for Biologists, Toxicologists and Toxinologists, Physicians and Pharmacists. 1st edit Medpharm; 2002. [Google Scholar]
- 3.Dutertre S, Lewis RJ. Use of venom peptides to probe ion channel structure and function. J Biol Chem. 2010;285:13315–20. doi: 10.1074/jbc.R109.076596. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Bohlen CJ, Priel A, Zhou S, King D, Siemens J, Julius D. A bivalent tarantula toxin activates the capsaicin receptor, TRPV1, by targeting the outer pore domain. Cell. 2010;141:834–45. doi: 10.1016/j.cell.2010.03.052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Cao E, Cordero-Morales JF, Liu B, Qin F, Julius D. TRPV1 channels are intrinsically heat sensitive and negatively regulated by phosphoinositide lipids. Neuron. 2013;77:667–79. doi: 10.1016/j.neuron.2012.12.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Cao E, Liao M, Cheng Y, Julius D. TRPV1 structures in distinct conformations reveal activation mechanisms. Nature. 2013;504:113–8. doi: 10.1038/nature12823. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Diochot S, Salinas M, Baron A, Escoubas P, Lazdunski M. Peptides inhibitors of acid-sensing ion channels. Toxicon. 2007;49:271–284. doi: 10.1016/j.toxicon.2006.09.026. [DOI] [PubMed] [Google Scholar]
- 8.Baconguis I, Bohlen CJ, Goehring A, Julius D, Gouaux E. X-Ray Structure of Acid-Sensing Ion Channel 1-Snake Toxin Complex Reveals Open State of a Na+-Selective Channel. Cell. 2014;156:717–729. doi: 10.1016/j.cell.2014.01.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Bowman CL, Gottlieb PA, Suchyna TM, Murphy YK, Sachs F. Mechanosensitive ion channels and the peptide inhibitor GsMTx-4: History, properties, mechanisms and pharmacology. Toxicon. 2007;49:249–270. doi: 10.1016/j.toxicon.2006.09.030. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Debin JA, Strichartz GR. Chloride Channel Inhibition by the Venom of the Scorpion Leiurus-Quinquestriatus. Toxicon. 1991;29:1403–1408. doi: 10.1016/0041-0101(91)90128-e. [DOI] [PubMed] [Google Scholar]
- 11.Chang CC, Lee CY. Isolation of Neurotoxins from Venom of Bungarus Multicinctus and Their Modes of Neuromuscular Blocking Action. Archives Internationales De Pharmacodynamie Et De Therapie. 1963;144:241–&. [PubMed] [Google Scholar]
- 12.Platt RJ, Curtice KJ, Twede VD, Watkins M, Gruszczynski P, Bulaj G, Horvath MP, Olivera B. From molecular phylogeny towards differentiating pharmacology for NMDA receptor subtypes. Toxicon. 2014;81:67–79. doi: 10.1016/j.toxicon.2014.01.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Maiga A, Mourier G, Quinton L, Rouget C, Gales C, Denis C, Lluel P, Senard JM, Palea S, Servent D, Gilles N. G protein-coupled receptors, an unexploited animal toxin targets: Exploration of green mamba venom for novel drug candidates active against adrenoceptors. Toxicon. 2012;59:487–96. doi: 10.1016/j.toxicon.2011.03.009. [DOI] [PubMed] [Google Scholar]
- 14.Habermann E, Ahnert-Hilger G, Chhatwal GS, Beress L. Delayed haemolytic action of palytoxin. General characteristics. Biochim Biophys Acta. 1981;649:481–6. doi: 10.1016/0005-2736(81)90439-9. [DOI] [PubMed] [Google Scholar]
- 15.Hille B. The receptor for tetrodotoxin and saxitoxin. A structural hypothesis. Biophys J. 1975;15:615–619. doi: 10.1016/S0006-3495(75)85842-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Olivera BM. Conus venom peptides, receptor and ion channel targets and drug design: 50 million years of neuropharmacology (E.E. Just Lecture, 1996). Mol Biol Cell. 1997;8:2101–2109. doi: 10.1091/mbc.8.11.2101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Cahalan MD. Modification of sodium channel gating in frog myelinated nerve fibres by Centruroides sculpturatus scorpion venom. J Physiol. 1975;244:511–34. doi: 10.1113/jphysiol.1975.sp010810. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Koppenhofer E, Schmidt H. [Effect of scorpion venom on ionic currents of the node of Ranvier. II. Incomplete sodium inactivation]. Pflugers Arch. 1968;303:150–61. doi: 10.1007/BF00592632. [DOI] [PubMed] [Google Scholar]
- 19.Koppenhofer E, Schmidt H. [Effect of scorpion venom on ionic currents of the node of Ranvier. I. The permeabilities PNa and PK]. Pflugers Arch. 1968;303:133–49. doi: 10.1007/BF00592631. [DOI] [PubMed] [Google Scholar]
- 20.Alabi AA, Bahamonde MI, Jung HJ, Kim JI, Swartz KJ. Portability of paddle motif function and pharmacology in voltage sensors. Nature. 2007;450:370–5. doi: 10.1038/nature06266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Chakrapani S, Cuello LG, Cortes DM, Perozo E. Structural dynamics of an isolated voltage-sensor domain in a lipid bilayer. Structure. 2008;16:398–409. doi: 10.1016/j.str.2007.12.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Jiang Y, Lee A, Chen J, Ruta V, Cadene M, Chait BT, MacKinnon R. X-ray structure of a voltage-dependent K+ channel. Nature. 2003;423:33–41. doi: 10.1038/nature01580. [DOI] [PubMed] [Google Scholar]
- 23.Long SB, Tao X, Campbell EB, MacKinnon R. Atomic structure of a voltage-dependent K+ channel in a lipid membrane-like environment. Nature. 2007;450:376–82. doi: 10.1038/nature06265. [DOI] [PubMed] [Google Scholar]
- 24.Swartz KJ. Sensing voltage across lipid membranes. Nature. 2008;456:891–7. doi: 10.1038/nature07620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Swartz KJ. Tarantula toxins interacting with voltage sensors in potassium channels. Toxicon. 2007;49:213–30. doi: 10.1016/j.toxicon.2006.09.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Garcia ML, Hanner M, Knaus HG, Koch R, Schmalhofer W, Slaughter RS, Kaczorowski GJ. Pharmacology of potassium channels. Advances in pharmacology. 1997;39:425–71. doi: 10.1016/s1054-3589(08)60078-2. [DOI] [PubMed] [Google Scholar]
- 27.MacKinnon R. Determination of the subunit stoichiometry of a voltage-activated potassium channel. Nature. 1991;350:232–5. doi: 10.1038/350232a0. [DOI] [PubMed] [Google Scholar]
- 28.Hidalgo P, MacKinnon R. Revealing the architecture of a K+ channel pore through mutant cycles with a peptide inhibitor. Science. 1995;268:307–10. doi: 10.1126/science.7716527. [DOI] [PubMed] [Google Scholar]
- 29.Diochot S, Lazdunski M. Sea anemone toxins affecting potassium channels. Progress in molecular and subcellular biology. 2009;46:99–122. doi: 10.1007/978-3-540-87895-7_4. [DOI] [PubMed] [Google Scholar]
- 30.Harvey AL. Twenty years of dendrotoxins. Toxicon : official journal of the International Society on Toxinology. 2001;39:15–26. doi: 10.1016/s0041-0101(00)00162-8. [DOI] [PubMed] [Google Scholar]
- 31.Weatherall KL, Goodchild SJ, Jane DE, Marrion NV. Small conductance calcium-activated potassium channels: from structure to function. Progress in neurobiology. 2010;91:242–55. doi: 10.1016/j.pneurobio.2010.03.002. [DOI] [PubMed] [Google Scholar]
- 32.Honma T, Shiomi K. Peptide toxins in sea anemones: structural and functional aspects. Marine biotechnology. 2006;8:1–10. doi: 10.1007/s10126-005-5093-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 33.Miller C, Moczydlowski E, Latorre R, Phillips M. Charybdotoxin, a protein inhibitor of single Ca2+-activated K+ channels from mammalian skeletal muscle. Nature. 1985;313:316–8. doi: 10.1038/313316a0. [DOI] [PubMed] [Google Scholar]
- 34.Anderson CS, MacKinnon R, Smith C, Miller C. Charybdotoxin block of single Ca2+-activated K+ channels. Effects of channel gating, voltage, and ionic strength. J Gen Physiol. 1988;91:317–33. doi: 10.1085/jgp.91.3.317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.MacKinnon R, Miller C. Mechanism of charybdotoxin block of the high-conductance, Ca2+-activated K+ channel. J Gen Physiol. 1988;91:335–49. doi: 10.1085/jgp.91.3.335. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 36.Park CS, Miller C. Interaction of charybdotoxin with permeant ions inside the pore of a K+ channel. Neuron. 1992;9:307–13. doi: 10.1016/0896-6273(92)90169-e. [DOI] [PubMed] [Google Scholar]
- 37.Garcia ML, Garcia-Calvo M, Hidalgo P, Lee A, MacKinnon R. Purification and characterization of three inhibitors of voltage-dependent K+ channels from Leiurus quinquestriatus var. hebraeus venom. Biochemistry. 1994;33:6834–9. doi: 10.1021/bi00188a012. [DOI] [PubMed] [Google Scholar]
- 38.Miller C. The charybdotoxin family of K+ channel-blocking peptides. Neuron. 1995;15:5–10. doi: 10.1016/0896-6273(95)90057-8. [DOI] [PubMed] [Google Scholar]
- 39.Banerjee A, Lee A, Campbell E, Mackinnon R. Structure of a pore-blocking toxin in complex with a eukaryotic voltage-dependent K(+) channel. eLife. 2013;2:e00594. doi: 10.7554/eLife.00594. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Lange A, Giller K, Hornig S, Martin-Eauclaire MF, Pongs O, Becker S, Baldus M. Toxin-induced conformational changes in a potassium channel revealed by solid-state NMR. Nature. 2006;440:959–62. doi: 10.1038/nature04649. [DOI] [PubMed] [Google Scholar]
- 41.Swartz KJ, MacKinnon R. Mapping the receptor site for hanatoxin, a gating modifier of voltage-dependent K+ channels. Neuron. 1997;18:675–82. doi: 10.1016/s0896-6273(00)80307-4. [DOI] [PubMed] [Google Scholar]
- 42.Swartz KJ, MacKinnon R. Hanatoxin modifies the gating of a voltage-dependent K+ channel through multiple binding sites. Neuron. 1997;18:665–73. doi: 10.1016/s0896-6273(00)80306-2. [DOI] [PubMed] [Google Scholar]
- 43.Phillips LR, Milescu M, Li-Smerin Y, Mindell JA, Kim JI, Swartz KJ. Voltage-sensor activation with a tarantula toxin as cargo. Nature. 2005;436:857–60. doi: 10.1038/nature03873. [DOI] [PubMed] [Google Scholar]
- 44.Bezanilla F, Stefani E. Gating currents. Methods Enzymol. 1998;293:331–52. doi: 10.1016/s0076-6879(98)93022-1. [DOI] [PubMed] [Google Scholar]
- 45.Lee HC, Wang JM, Swartz KJ. Interaction between extracellular Hanatoxin and the resting conformation of the voltage-sensor paddle in Kv channels. Neuron. 2003;40:527–36. doi: 10.1016/s0896-6273(03)00636-6. [DOI] [PubMed] [Google Scholar]
- 46.Li-Smerin Y, Swartz KJ. Gating modifier toxins reveal a conserved structural motif in voltage-gated Ca2+ and K+ channels. Proc Natl Acad Sci U S A. 1998;95:8585–9. doi: 10.1073/pnas.95.15.8585. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Li-Smerin Y, Swartz KJ. Localization and molecular determinants of the Hanatoxin receptors on the voltage-sensing domains of a K(+) channel. J Gen Physiol. 2000;115:673–84. doi: 10.1085/jgp.115.6.673. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Ruta V, MacKinnon R. Localization of the voltage-sensor toxin receptor on KvAP. Biochemistry. 2004;43:10071–9. doi: 10.1021/bi049463y. [DOI] [PubMed] [Google Scholar]
- 49.Jung HJ, Lee JY, Kim SH, Eu YJ, Shin SY, Milescu M, Swartz KJ, Kim JI. Solution structure and lipid membrane partitioning of VSTx1, an inhibitor of the KvAP potassium channel. Biochemistry. 2005;44:6015–23. doi: 10.1021/bi0477034. [DOI] [PubMed] [Google Scholar]
- 50.Milescu M, Bosmans F, Lee S, Alabi AA, Kim JI, Swartz KJ. Interactions between lipids and voltage sensor paddles detected with tarantula toxins. Nat Struct Mol Biol. 2009;16:1080–5. doi: 10.1038/nsmb.1679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Milescu M, Vobecky J, Roh SH, Kim SH, Jung HJ, Kim JI, Swartz KJ. Tarantula toxins interact with voltage sensors within lipid membranes. J Gen Physiol. 2007;130:497–511. doi: 10.1085/jgp.200709869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 52.Takahashi H, Kim JI, Min HJ, Sato K, Swartz KJ, Shimada I. Solution structure of hanatoxin1, a gating modifier of voltage-dependent K(+) channels: common surface features of gating modifier toxins. J Mol Biol. 2000;297:771–80. doi: 10.1006/jmbi.2000.3609. [DOI] [PubMed] [Google Scholar]
- 53.Wang JM, Roh SH, Kim S, Lee CW, Kim JI, Swartz KJ. Molecular surface of tarantula toxins interacting with voltage sensors in K(v) channels. J Gen Physiol. 2004;123:455–67. doi: 10.1085/jgp.200309005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Bosmans F, Martin-Eauclaire MF, Swartz KJ. Deconstructing voltage sensor function and pharmacology in sodium channels. Nature. 2008;456:202–8. doi: 10.1038/nature07473. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.Middleton RE, Warren VA, Kraus RL, Hwang JC, Liu CJ, Dai G, Brochu RM, Kohler MG, Gao YD, Garsky VM, Bogusky MJ, Mehl JT, Cohen CJ, Smith MM. Two tarantula peptides inhibit activation of multiple sodium channels. Biochemistry. 2002;41:14734–47. doi: 10.1021/bi026546a. [DOI] [PubMed] [Google Scholar]
- 56.Edgerton GB, Blumenthal K, Hanck D. Modification of gating kinetics in Cav3.1 by the tarantula toxin ProTxII. Biophys J. 2007:601a–601a. [Google Scholar]
- 57.Liu P, Jo S, Bean BP. Modulation of neuronal sodium channels by the sea anemone peptide BDS-I. Journal of neurophysiology. 2012;107:3155–67. doi: 10.1152/jn.00785.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Goldin AL, Barchi RL, Caldwell JH, Hofmann F, Howe JR, Hunter JC, Kallen RG, Mandel G, Meisler MH, Netter YB, Noda M, Tamkun MM, Waxman SG, Wood JN, Catterall WA. Nomenclature of voltage-gated sodium channels. Neuron. 2000;28:365–8. doi: 10.1016/s0896-6273(00)00116-1. [DOI] [PubMed] [Google Scholar]
- 59.Campos FV, Chanda B, Beirao PS, Bezanilla F. Alpha-scorpion toxin impairs a conformational change that leads to fast inactivation of muscle sodium channels. J Gen Physiol. 2008;132:251–63. doi: 10.1085/jgp.200809995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 60.Cha A, Ruben PC, George AL, Jr., Fujimoto E, Bezanilla F. Voltage sensors in domains III and IV, but not I and II, are immobilized by Na+ channel fast inactivation. Neuron. 1999;22:73–87. doi: 10.1016/s0896-6273(00)80680-7. [DOI] [PubMed] [Google Scholar]
- 61.Chanda B, Bezanilla F. Tracking voltage-dependent conformational changes in skeletal muscle sodium channel during activation. J Gen Physiol. 2002;120:629–45. doi: 10.1085/jgp.20028679. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Horn R, Ding S, Gruber HJ. Immobilizing the moving parts of voltage-gated ion channels. J Gen Physiol. 2000;116:461–76. doi: 10.1085/jgp.116.3.461. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 63.Sheets MF, Kyle JW, Hanck DA. The role of the putative inactivation lid in sodium channel gating current immobilization. J Gen Physiol. 2000;115:609–20. doi: 10.1085/jgp.115.5.609. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Sheets MF, Kyle JW, Kallen RG, Hanck DA. The Na channel voltage sensor associated with inactivation is localized to the external charged residues of domain IV, S4. Biophys J. 1999;77:747–57. doi: 10.1016/S0006-3495(99)76929-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Payandeh J, Gamal El-Din TM, Scheuer T, Zheng N, Catterall WA. Crystal structure of a voltage-gated sodium channel in two potentially inactivated states. Nature. 2012;486:135–9. doi: 10.1038/nature11077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Payandeh J, Scheuer T, Zheng N, Catterall WA. The crystal structure of a voltage-gated sodium channel. Nature. 2011;475:353–8. doi: 10.1038/nature10238. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Shaya D, Findeisen F, Abderemane-Ali F, Arrigoni C, Wong S, Nurva SR, Loussouarn G, Minor DL., Jr. Structure of a prokaryotic sodium channel pore reveals essential gating elements and an outer ion binding site common to eukaryotic channels. J Mol Biol. 2014;426:467–83. doi: 10.1016/j.jmb.2013.10.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Zhang X, Ren W, DeCaen P, Yan C, Tao X, Tang L, Wang J, Hasegawa K, Kumasaka T, He J, Clapham DE, Yan N. Crystal structure of an orthologue of the NaChBac voltage-gated sodium channel. Nature. 2012;486:130–4. doi: 10.1038/nature11054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Hodgkin AL, Huxley AF. The dual effect of membrane potential on sodium conductance in the giant axon of Loligo. J Physiol. 1952;116:497–506. doi: 10.1113/jphysiol.1952.sp004719. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Hodgkin AL, Huxley AF. Currents carried by sodium and potassium ions through the membrane of the giant axon of Loligo. J Physiol. 1952;116:449–72. doi: 10.1113/jphysiol.1952.sp004717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Furukawa T, Sasaoka T, Hosoya Y. Effects of tetrodotoxin on the neuromuscular junction. Japanese J Phys. 1959;9:143–52. doi: 10.2170/jjphysiol.9.143. [DOI] [PubMed] [Google Scholar]
- 72.Narahashi T, Anderson N, Moore J. Comparison of tetrodotoxin and procaine in internally perfused squid giant axons. J Gen Physiol. 1967;50:1413–1428. doi: 10.1085/jgp.50.5.1413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Narahashi T, Moore JW, Scott WR. Tetrodotoxin Blockage of Sodium Conductance Increase in Lobster Giant Axons. J Gen Physiol. 1964;47:965–74. doi: 10.1085/jgp.47.5.965. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Woodward R. The structure of tetrodotoxin. Pure Appl. Chem. 1964;9:49–74. [Google Scholar]
- 75.Agnew WS, Levinson SR, Brabson JS, Raftery MA. Purification of the tetrodotoxin-binding component associated with the voltage-sensitive sodium channel from Electrophorus electricus electroplax membranes. Proc Natl Acad Sci U S A. 1978;75:2606–10. doi: 10.1073/pnas.75.6.2606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Miller J, Agnew W, Levinson S. Principal glycopeptide of the tetrodotoxin/saxitoxin binding protein from Electrophorus electricus: isolation and partial chemical and physical characterization. Biochemistry. 1983;22:462–470. doi: 10.1021/bi00271a032. [DOI] [PubMed] [Google Scholar]
- 77.Barchi R, Cohen S, Murphy L. Purification from rat sarcolemma of the saxitoxin-binding component of the excitable membrane sodium channel. Proc Natl Acad Sci U S A. 1980;77:1306–1310. doi: 10.1073/pnas.77.3.1306. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Hartshorne R, Catterall W. Purification of the saxitoxin receptor of the sodium channel from rat brain. Proc Natl Acad Sci U S A. 1981;78:4620–4624. doi: 10.1073/pnas.78.7.4620. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Hartshorne R, Messner D, Coppersmith J, Catterall W. The saxitoxin receptor of the sodium channel from rat brain. Evidence for two nonidentical beta subunits. J Biol Chem. 1982;257:13888–13891. [PubMed] [Google Scholar]
- 80.Hoehne A, Behera D, Parsons WH, James ML, Shen B, Borgohain P, Bodapati D, Prabhakar A, Gambhir SS, Yeomans DC, Biswal S, Chin FT, Du Bois J. A 18F-labeled saxitoxin derivative for in vivo PET-MR imaging of voltage-gated sodium channel expression following nerve injury. J Am Chem Soc. 2013;135:18012–5. doi: 10.1021/ja408300e. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Ondrus AE, Lee HL, Iwanaga S, Parsons WH, Andresen BM, Moerner WE, Du Bois J. Fluorescent saxitoxins for live cell imaging of single voltage-gated sodium ion channels beyond the optical diffraction limit. Chemistry & biology. 2012;19:902–12. doi: 10.1016/j.chembiol.2012.05.021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Leffler A, Herzog RI, Dib-Hajj SD, Waxman SG, Cummins TR. Pharmacological properties of neuronal TTX-resistant sodium channels and the role of a critical serine pore residue. Pflugers Archiv. 2005;451:454–63. doi: 10.1007/s00424-005-1463-x. [DOI] [PubMed] [Google Scholar]
- 83.Lipkind G, Fozzard H. A structural model of the tetrodotoxin and saxitoxin binding site of the Na+ channel. Biophys J. 1994;66:1–13. doi: 10.1016/S0006-3495(94)80746-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Sivilotti L, Okuse K, Akopian AN, Moss S, Wood JN. A single serine residue confers tetrodotoxin insensitivity on the rat sensory-neuron-specific sodium channel SNS. FEBS letters. 1997;409:49–52. doi: 10.1016/s0014-5793(97)00479-1. [DOI] [PubMed] [Google Scholar]
- 85.Santarelli VP, Eastwood AL, Dougherty DA, Horn R, Ahern CA. A cation-pi interaction discriminates among sodium channels that are either sensitive or resistant to tetrodotoxin block. J Biol Chem. 2007;282:8044–51. doi: 10.1074/jbc.M611334200. [DOI] [PubMed] [Google Scholar]
- 86.Lipkind GM, Fozzard HA. Voltage-gated Na channel selectivity: the role of the conserved domain III lysine residue. J Gen Physiol. 2008;131:523–9. doi: 10.1085/jgp.200809991. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Terlau H, Heinemann S, Stühmer W, Pusch M, Conti F, Imoto K, Numa S. Mapping the site of block by tetrodotoxin and saxitoxin of sodium channel II. FEBS letters. 1991;293:93–96. doi: 10.1016/0014-5793(91)81159-6. [DOI] [PubMed] [Google Scholar]
- 88.Martin-Eauclaire MF, Couraud F. Scorpion neurotoxins: effects and mechanisms. In: Chang LW, Dyer RS, editors. Handk. Neurotoxicology. Marcel-Dekker; New York: 1992. [Google Scholar]
- 89.Jover E, Martin-Moutot N, Couraud F, Rochat H. Scorpion toxin: specific binding to rat synaptosomes. Biochem Biophys Res Comm. 1978;85:377–382. doi: 10.1016/s0006-291x(78)80053-9. [DOI] [PubMed] [Google Scholar]
- 90.Couraud F, Jover E, Dubois J, Rochat H. Two types of scorpion receptor sites, one related to the activation, the other to the inactivation of the action potential sodium channel. Toxicon. 1982;20:9–16. doi: 10.1016/0041-0101(82)90138-6. [DOI] [PubMed] [Google Scholar]
- 91.Tejedor F, Catterall W. Site of covalent attachment of alpha-scorpion toxin derivatives in domain I of the sodium channel alpha subunit. Proc Natl Acad Sci U S A. 1988;85:8742–8746. doi: 10.1073/pnas.85.22.8742. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Thomsen W, Catterall W. Localization of the receptor site for alpha-scorpion toxins by antibody mapping: implications for sodium channel topology. Proc Natl Acad Sci U S A. 1989;86:10161–10165. doi: 10.1073/pnas.86.24.10161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 93.Rogers J, Qu Y, Tanada T, Scheuer T, Catterall W. 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. 1996;271:15950–15962. doi: 10.1074/jbc.271.27.15950. [DOI] [PubMed] [Google Scholar]
- 94.Wang J, Yarov-Yarovoy V, Kahn R, Gordon D, Gurevitz M, Scheuer T, Catterall W. Mapping the receptor site for alpha-scorpion toxins on a Na+ channel voltage sensor. Proc Natl Acad Sci U S A. 2011;108:15426–15431. doi: 10.1073/pnas.1112320108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Campos F, Coronas F, Beirão P. Voltage-dependent displacement of the scorpion toxin Ts3 from sodium channels and its implication on the control of inactivation. British journal of pharmacology. 2004;142:1115–1122. doi: 10.1038/sj.bjp.0705793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Leipold E, Borges A, Heinemann SH. Scorpion beta-toxin interference with NaV channel voltage sensor gives rise to excitatory and depressant modes. J Gen Physiol. 2012;139:305–19. doi: 10.1085/jgp.201110720. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Marcotte P, Chen LQ, Kallen RG, Chahine M. Effects of Tityus serrulatus scorpion toxin gamma on voltage-gated Na+ channels. Circ Res. 1997;80:363–9. doi: 10.1161/01.res.80.3.363. [DOI] [PubMed] [Google Scholar]
- 98.Cestèle S, Qu Y, Rogers J, Rochat H, Scheuer T, Catterall W. Voltage sensor-trapping: enhanced activation of sodium channels by beta-scorpion toxin bound to the S3-S4 loop in domain II. Neuron. 1998;21:919–931. doi: 10.1016/s0896-6273(00)80606-6. [DOI] [PubMed] [Google Scholar]
- 99.Cestèle S, Yarov-Yarovoy V, Qu Y, Sampieri F, Scheuer T, Catterall W. Structure and function of the voltage sensor of sodium channels probed by a beta-scorpion toxin. J Biol Chem. 2006;281:21332–21344. doi: 10.1074/jbc.M603814200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Bosmans F, Swartz KJ. Targeting voltage sensors in sodium channels with spider toxins. Trends in pharmacological sciences. 2010;31:175–182. doi: 10.1016/j.tips.2009.12.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Wang M, Diao J, Li J, Tang J, Lin Y, Hu W, Zhang Y, Xiao Y, Liang S. JZTX-IV, a unique acidic sodium channel toxin isolated from the spider Chilobrachys jingzhao. Toxicon. 2008;52:871–80. doi: 10.1016/j.toxicon.2008.08.018. [DOI] [PubMed] [Google Scholar]
- 102.Corzo G, Sabo JK, Bosmans F, Billen B, Villegas E, Tytgat J, Norton RS. Solution structure and alanine scan of a spider toxin that affects the activation of mammalian voltage-gated sodium channels. J Biol Chem. 2007;282:4643–52. doi: 10.1074/jbc.M605403200. [DOI] [PubMed] [Google Scholar]
- 103.Bosmans F, Puopolo M, Martin-Eauclaire MF, Bean BP, Swartz KJ. Functional properties and toxin pharmacology of a dorsal root ganglion sodium channel viewed through its voltage sensors. J Gen Physiol. 2011;138:59–72. doi: 10.1085/jgp.201110614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Schmalhofer WA, Calhoun J, Burrows R, Bailey T, Kohler MG, Weinglass AB, Kaczorowski GJ, Garcia ML, Koltzenburg M, Priest BT. ProTx-II, a selective inhibitor of NaV1.7 sodium channels, blocks action potential propagation in nociceptors. Mol Pharmacol. 2008;74:1476–84. doi: 10.1124/mol.108.047670. [DOI] [PubMed] [Google Scholar]
- 105.Xiao Y, Bingham JP, Zhu W, Moczydlowski E, Liang S, Cummins TR. Tarantula huwentoxin-IV inhibits neuronal sodium channels by binding to receptor site 4 and trapping the domain ii voltage sensor in the closed configuration. J Biol Chem. 2008;283:27300–13. doi: 10.1074/jbc.M708447200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Xiao Y, Blumenthal K, Jackson JO, 2nd, Liang S, Cummins TR. The tarantula toxins ProTx-II and huwentoxin-IV differentially interact with human Nav1.7 voltage sensors to inhibit channel activation and inactivation. Mol Pharmacol. 2010;78:1124–34. doi: 10.1124/mol.110.066332. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Xiao Y, Jackson JO, 2nd, Liang S, Cummins TR. Common molecular determinants of tarantula huwentoxin-IV inhibition of Na+ channel voltage sensors in domains II and IV. J Biol Chem. 2011;286:27301–10. doi: 10.1074/jbc.M111.246876. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Catterall WA. Voltage-gated calcium channels. Cold Spring Harb Perspect Biol. 2011;3:a003947. doi: 10.1101/cshperspect.a003947. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Tsien RW, Lipscombe D, Madison D, Bley K, Fox A. Reflections on Ca(2+)-channel diversity, 1988-1994. Trends in neurosciences. 1995;18:52–4. [PubMed] [Google Scholar]
- 110.Dolphin AC. Calcium channel diversity: multiple roles of calcium channel subunits. Current opinion in neurobiology. 2009;19:237–44. doi: 10.1016/j.conb.2009.06.006. [DOI] [PubMed] [Google Scholar]
- 111.Simms BA, Zamponi GW. Neuronal voltage-gated calcium channels: structure, function, and dysfunction. Neuron. 2014;82:24–45. doi: 10.1016/j.neuron.2014.03.016. [DOI] [PubMed] [Google Scholar]
- 112.Catterall WA, Perez-Reyes E, Snutch TP, Striessnig J. International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharmacol Rev. 2005;57:411–25. doi: 10.1124/pr.57.4.5. [DOI] [PubMed] [Google Scholar]
- 113.Hofmann F, Flockerzi V, Kahl S, Wegener JW. L-type CaV1.2 calcium channels: from in vitro findings to in vivo function. Physiol rev. 2014;94:303–26. doi: 10.1152/physrev.00016.2013. [DOI] [PubMed] [Google Scholar]
- 114.Cheong E, Shin HS. T-type Ca2+ channels in normal and abnormal brain functions. Physiol rev. 2013;93:961–92. doi: 10.1152/physrev.00010.2012. [DOI] [PubMed] [Google Scholar]
- 115.Lipscombe D, Allen SE, Toro CP. Control of neuronal voltage-gated calcium ion channels from RNA to protein. Trends in Neurosciences. 2013;36:598–609. doi: 10.1016/j.tins.2013.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 116.Ferry DR, Glossmann H. Identification of putative calcium channels in skeletal muscle microsomes. FEBS letters. 1982;148:331–7. doi: 10.1016/0014-5793(82)80835-1. [DOI] [PubMed] [Google Scholar]
- 117.Reuter H. Calcium channel modulation by neurotransmitters, enzymes and drugs. Nature. 1983;301:569–74. doi: 10.1038/301569a0. [DOI] [PubMed] [Google Scholar]
- 118.Flockerzi V, Oeken HJ, Hofmann F, Pelzer D, Cavalie A, Trautwein W. Purified dihydropyridine-binding site from skeletal muscle t-tubules is a functional calcium channel. Nature. 1986;323:66–8. doi: 10.1038/323066a0. [DOI] [PubMed] [Google Scholar]
- 119.McCleskey EW, Fox AP, Feldman DH, Cruz LJ, Olivera BM, Tsien RW, Yoshikami D. Omega-conotoxin: direct and persistent blockade of specific types of calcium channels in neurons but not muscle. Proc Natl Acad Sci U S A. 1987;84:4327–31. doi: 10.1073/pnas.84.12.4327. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 120.Olivera BM, McIntosh JM, Cruz LJ, Luque FA, Gray WR. Purification and sequence of a presynaptic peptide toxin from Conus geographus venom. Biochemistry. 1984;23:5087–90. doi: 10.1021/bi00317a001. [DOI] [PubMed] [Google Scholar]
- 121.Olivera BM, Cruz LJ, de Santos V, LeCheminant GW, Griffin D, Zeikus R, McIntosh JM, Galyean R, Varga J, Gray WR, et al. Neuronal calcium channel antagonists. Discrimination between calcium channel subtypes using omega-conotoxin from Conus magus venom. Biochemistry. 1987;26:2086–90. doi: 10.1021/bi00382a004. [DOI] [PubMed] [Google Scholar]
- 122.Cruz LJ, Olivera BM. Calcium channel antagonists. Omega-conotoxin defines a new high affinity site. J Biol Chem. 1986;261:6230–3. [PubMed] [Google Scholar]
- 123.Williams ME, Brust PF, Feldman DH, Patthi S, Simerson S, Maroufi A, McCue AF, Velicelebi G, Ellis SB, Harpold MM. Structure and functional expression of an omega-conotoxin-sensitive human N-type calcium channel. Science. 1992;257:389–95. doi: 10.1126/science.1321501. [DOI] [PubMed] [Google Scholar]
- 124.Williams ME, Feldman DH, McCue AF, Brenner R, Velicelebi G, Ellis SB, Harpold MM. Structure and functional expression of alpha 1, alpha 2, and beta subunits of a novel human neuronal calcium channel subtype. Neuron. 1992;8:71–84. doi: 10.1016/0896-6273(92)90109-q. [DOI] [PubMed] [Google Scholar]
- 125.Ellinor PT, Zhang JF, Horne WA, Tsien RW. Structural determinants of the blockade of N-type calcium channels by a peptide neurotoxin. Nature. 1994;372:272–5. doi: 10.1038/372272a0. [DOI] [PubMed] [Google Scholar]
- 126.Feng ZP, Hamid J, Doering C, Bosey GM, Snutch TP, Zamponi GW. Residue Gly1326 of the N-type calcium channel alpha 1B subunit controls reversibility of omega-conotoxin GVIA and MVIIA block. The Journal of biological chemistry. 2001;276:15728–35. doi: 10.1074/jbc.M100406200. [DOI] [PubMed] [Google Scholar]
- 127.Stocker JW, Nadasdi L, Aldrich RW, Tsien RW. Preferential interaction of omega-conotoxins with inactivated N-type Ca2+ channels. J Neurosci. 1997;17:3002–13. doi: 10.1523/JNEUROSCI.17-09-03002.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Boland LM, Morrill JA, Bean BP. omega-Conotoxin block of N-type calcium channels in frog and rat sympathetic neurons. J Neurosci. 1994;14:5011–27. doi: 10.1523/JNEUROSCI.14-08-05011.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 129.Liang H, Elmslie KS. Rapid and reversible block of N-type calcium channels (CaV 2.2) by omega-conotoxin GVIA in the absence of divalent cations. J Neurosci. 2002;22:8884–90. doi: 10.1523/JNEUROSCI.22-20-08884.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 130.McDonough SI, Boland LM, Mintz IM, Bean BP. Interactions among toxins that inhibit N-type and P-type calcium channels. The Journal of general physiology. 2002;119:313–28. doi: 10.1085/jgp.20028560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 131.Jones LP, DeMaria CD, Yue DT. N-type calcium channel inactivation probed by gating-current analysis. Biophysical journal. 1999;76:2530–52. doi: 10.1016/S0006-3495(99)77407-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Yarotskyy V, Elmslie KS. omega-Conotoxin GVIA Alters Gating Charge Movement of N-Type (CaV2.2) Calcium Channels. Journal of neurophysiology. 2009;101:332–340. doi: 10.1152/jn.91064.2008. [DOI] [PubMed] [Google Scholar]
- 133.Pringos E, Vignes M, Martinez J, Rolland V. Peptide neurotoxins that affect voltage-gated calcium channels: a close-up on omega-agatoxins. Toxins. 2011;3:17–42. doi: 10.3390/toxins3010017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.Llinas R, Yarom Y. Properties and distribution of ionic conductances generating electroresponsiveness of mammalian inferior olivary neurones in vitro. J Physiol. 1981;315:569–84. doi: 10.1113/jphysiol.1981.sp013764. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 135.Llinas RR, Sugimori M, Cherksey B. Voltage-dependent calcium conductances in mammalian neurons. The P channel. Ann N Y Acad Sci. 1989;560:103–11. doi: 10.1111/j.1749-6632.1989.tb24084.x. [DOI] [PubMed] [Google Scholar]
- 136.Hillyard DR, Monje VD, Mintz IM, Bean BP, Nadasdi L, Ramachandran J, Miljanich G, Azimi-Zoonooz A, McIntosh JM, Cruz LJ, et al. A new Conus peptide ligand for mammalian presynaptic Ca2+ channels. Neuron. 1992;9:69–77. doi: 10.1016/0896-6273(92)90221-x. [DOI] [PubMed] [Google Scholar]
- 137.Mintz IM, Adams ME, Bean BP. P-type calcium channels in rat central and peripheral neurons. Neuron. 1992;9:85–95. doi: 10.1016/0896-6273(92)90223-z. [DOI] [PubMed] [Google Scholar]
- 138.Mintz IM, Venema VJ, Swiderek KM, Lee TD, Bean BP, Adams ME. P-type calcium channels blocked by the spider toxin omega-Aga-IVA. Nature. 1992;355:827–9. doi: 10.1038/355827a0. [DOI] [PubMed] [Google Scholar]
- 139.Randall A, Tsien RW. Pharmacological dissection of multiple types of Ca2+ channel currents in rat cerebellar granule neurons. J Neurosci. 1995;15:2995–3012. doi: 10.1523/JNEUROSCI.15-04-02995.1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.McEnery MW, Snowman AM, Sharp AH, Adams ME, Snyder SH. Purified omega-conotoxin GVIA receptor of rat brain resembles a dihydropyridine-sensitive L-type calcium channel. Proc Natl Acad Sci U S A. 1991;88:11095–9. doi: 10.1073/pnas.88.24.11095. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 141.Martin-Moutot N, Leveque C, Sato K, Kato R, Takahashi M, Seagar M. Properties of omega conotoxin MVIIC receptors associated with alpha 1A calcium channel subunits in rat brain. FEBS Lett. 1995;366:21–5. doi: 10.1016/0014-5793(95)00467-n. [DOI] [PubMed] [Google Scholar]
- 142.Witcher DR, De Waard M, Liu H, Pragnell M, Campbell KP. Association of native Ca2+ channel beta subunits with the alpha 1 subunit interaction domain. J Biol Chem. 1995;270:18088–93. doi: 10.1074/jbc.270.30.18088. [DOI] [PubMed] [Google Scholar]
- 143.Lipscombe D, Pan JQ, Gray AC. Functional diversity in neuronal voltage-gated calcium channels by alternative splicing of Ca(v)alpha1. Mol Neurobiol. 2002;26:21–44. doi: 10.1385/MN:26:1:021. [DOI] [PubMed] [Google Scholar]
- 144.Bourinet E, Soong TW, Sutton K, Slaymaker S, Mathews E, Monteil A, Zamponi GW, Nargeot J, Snutch TP. Splicing of alpha 1A subunit gene generates phenotypic variants of P- and Q-type calcium channels. Nat Neurosci. 1999;2:407–15. doi: 10.1038/8070. [DOI] [PubMed] [Google Scholar]
- 145.Hans M, Urrutia A, Deal C, Brust PF, Stauderman K, Ellis SB, Harpold MM, Johnson EC, Williams ME. Structural elements in domain IV that influence biophysical and pharmacological properties of human alpha1A-containing high-voltage-activated calcium channels. Biophysical journal. 1999;76:1384–400. doi: 10.1016/S0006-3495(99)77300-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 146.Winterfield JR, Swartz KJ. A hot spot for the interaction of gating modifier toxins with voltage-dependent ion channels. The Journal of general physiology. 2000;116:637–44. doi: 10.1085/jgp.116.5.637. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 147.Tsunemi T, Saegusa H, Ishikawa K, Nagayama S, Murakoshi T, Mizusawa H, Tanabe T. Novel Cav2.1 splice variants isolated from Purkinje cells do not generate P-type Ca2+ current. J Biol Chem. 2002;277:7214–21. doi: 10.1074/jbc.M108222200. [DOI] [PubMed] [Google Scholar]
- 148.Mermelstein PG, Foehring RC, Tkatch T, Song WJ, Baranauskas G, Surmeier DJ. Properties of Q-type calcium channels in neostriatal and cortical neurons are correlated with beta subunit expression. J Neurosci. 1999;19:7268–77. doi: 10.1523/JNEUROSCI.19-17-07268.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 149.McDonough SI, Lampe RA, Keith RA, Bean BP. Voltage-dependent inhibition of N- and P-type calcium channels by the peptide toxin omega-grammotoxin-SIA. Mol Pharmacol. 1997;52:1095–104. doi: 10.1124/mol.52.6.1095. [DOI] [PubMed] [Google Scholar]
- 150.Bourinet E, Stotz SC, Spaetgens RL, Dayanithi G, Lemos J, Nargeot J, Zamponi GW. Interaction of SNX482 with domains III and IV inhibits activation gating of alpha(1E) (Ca(V)2.3) calcium channels. Biophys J. 2001;81:79–88. doi: 10.1016/S0006-3495(01)75681-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Chuang RS, Jaffe H, Cribbs L, Perez-Reyes E, Swartz KJ. Inhibition of T-type voltage-gated calcium channels by a new scorpion toxin. Nat Neurosci. 1998;1:668–74. doi: 10.1038/3669. [DOI] [PubMed] [Google Scholar]
- 152.Lee CW, Bae C, Lee J, Ryu JH, Kim HH, Kohno T, Swartz KJ, Kim JI. Solution structure of kurtoxin: a gating modifier selective for Cav3 voltage-gated Ca(2+) channels. Biochemistry. 2012;51:1862–73. doi: 10.1021/bi201633j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 153.Edgerton GB, Blumenthal KM, Hanck DA. Inhibition of the activation pathway of the T-type calcium channel Ca(V)3.1 by ProTxII. Toxicon. 2010;56:624–36. doi: 10.1016/j.toxicon.2010.06.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 154.Ohkubo T, Yamazaki J, Kitamura K. Tarantula toxin ProTx-I differentiates between human T-type voltage-gated Ca2+ Channels Cav3.1 and Cav3.2. J Pharmacol Sci. 2010;112:452–8. doi: 10.1254/jphs.09356fp. [DOI] [PubMed] [Google Scholar]
- 155.Venkatachalam K, Montell C. TRP channels. Annu Rev Biochem. 2007;76:387–417. doi: 10.1146/annurev.biochem.75.103004.142819. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Ramsey IS, Delling M, Clapham DE. An introduction to TRP channels. Annu Rev Physiol. 2006;68:619–47. doi: 10.1146/annurev.physiol.68.040204.100431. [DOI] [PubMed] [Google Scholar]
- 157.Nilius B, Talavera K, Owsianik G, Prenen J, Droogmans G, Voets T. Gating of TRP channels: a voltage connection? J Physiol. 2005;567:35–44. doi: 10.1113/jphysiol.2005.088377. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 158.Kalia J, Swartz KJ. Exploring structure-function relationships between TRP and Kv channels. Sci Rep. 2013;3:1523. doi: 10.1038/srep01523. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 159.Baez-Nieto D, Castillo JP, Dragicevic C, Alvarez O, Latorre R. Thermo-TRP channels: biophysics of polymodal receptors. Adv Exp Med Biol. 2011;704:469–90. doi: 10.1007/978-94-007-0265-3_26. [DOI] [PubMed] [Google Scholar]
- 160.Dhaka A, Viswanath V, Patapoutian A. Trp ion channels and temperature sensation. Annu Rev Neurosci. 2006;29:135–61. doi: 10.1146/annurev.neuro.29.051605.112958. [DOI] [PubMed] [Google Scholar]
- 161.Salat K, Moniczewski A, Librowski T. Transient receptor potential channels - emerging novel drug targets for the treatment of pain. Curr Med Chem. 2013;20:1409–36. doi: 10.2174/09298673113209990107. [DOI] [PubMed] [Google Scholar]
- 162.Julius D. TRP channels and pain. Annu Rev Cell Dev Biol. 2013;29:355–84. doi: 10.1146/annurev-cellbio-101011-155833. [DOI] [PubMed] [Google Scholar]
- 163.Caterina MJ, Schumacher MA, Tominaga M, Rosen TA, Levine JD, Julius D. The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature. 1997;389:816–24. doi: 10.1038/39807. [DOI] [PubMed] [Google Scholar]
- 164.Jordt SE, Julius D. Molecular basis for species-specific sensitivity to “hot” chili peppers. Cell. 2002;108:421–30. doi: 10.1016/s0092-8674(02)00637-2. [DOI] [PubMed] [Google Scholar]
- 165.Siemens J, Zhou S, Piskorowski R, Nikai T, Lumpkin EA, Basbaum AI, King D, Julius D. Spider toxins activate the capsaicin receptor to produce inflammatory pain. Nature. 2006;444:208–12. doi: 10.1038/nature05285. [DOI] [PubMed] [Google Scholar]
- 166.Bae C, Kalia J, Song I, Yu J, Kim HH, Swartz KJ, Kim JI. High yield production and refolding of the double-knot toxin, an activator of TRPV1 channels. PLoS One. 2012;7:e51516. doi: 10.1371/journal.pone.0051516. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Liao M, Cao E, Julius D, Cheng Y. Structure of the TRPV1 ion channel determined by electron cryo-microscopy. Nature. 2013;504:107–12. doi: 10.1038/nature12822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Bezanilla F. How membrane proteins sense voltage. Nat Rev Mol Cell Biol. 2008;9:323–32. doi: 10.1038/nrm2376. [DOI] [PubMed] [Google Scholar]
- 169.Nilius B, Appendino G, Owsianik G. The transient receptor potential channel TRPA1: from gene to pathophysiology. Pflugers Arch. 2012;464:425–58. doi: 10.1007/s00424-012-1158-z. [DOI] [PubMed] [Google Scholar]
- 170.Bautista DM, Pellegrino M, Tsunozaki M. TRPA1: A gatekeeper for inflammation. Annu Rev Physiol. 2013;75:181–200. doi: 10.1146/annurev-physiol-030212-183811. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 171.Bautista DM, Jordt SE, Nikai T, Tsuruda PR, Read AJ, Poblete J, Yamoah EN, Basbaum AI, Julius D. TRPA1 mediates the inflammatory actions of environmental irritants and proalgesic agents. Cell. 2006;124:1269–82. doi: 10.1016/j.cell.2006.02.023. [DOI] [PubMed] [Google Scholar]
- 172.Bautista DM, Movahed P, Hinman A, Axelsson HE, Sterner O, Hogestatt ED, Julius D, Jordt SE, Zygmunt PM. Pungent products from garlic activate the sensory ion channel TRPA1. Proc Natl Acad Sci U S A. 2005;102:12248–52. doi: 10.1073/pnas.0505356102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Bandell M, Story GM, Hwang SW, Viswanath V, Eid SR, Petrus MJ, Earley TJ, Patapoutian A. Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin. Neuron. 2004;41:849–57. doi: 10.1016/s0896-6273(04)00150-3. [DOI] [PubMed] [Google Scholar]
- 174.Macpherson LJ, Geierstanger BH, Viswanath V, Bandell M, Eid SR, Hwang S, Patapoutian A. The pungency of garlic: activation of TRPA1 and TRPV1 in response to allicin. Curr Biol. 2005;15:929–34. doi: 10.1016/j.cub.2005.04.018. [DOI] [PubMed] [Google Scholar]
- 175.Story GM, Peier AM, Reeve AJ, Eid SR, Mosbacher J, Hricik TR, Earley TJ, Hergarden AC, Andersson DA, Hwang SW, McIntyre P, Jegla T, Bevan S, Patapoutian A. ANKTM1, a TRP-like channel expressed in nociceptive neurons, is activated by cold temperatures. Cell. 2003;112:819–29. doi: 10.1016/s0092-8674(03)00158-2. [DOI] [PubMed] [Google Scholar]
- 176.Hill K, Schaefer M. TRPA1 is differentially modulated by the amphipathic molecules trinitrophenol and chlorpromazine. J Biol Chem. 2007;282:7145–53. doi: 10.1074/jbc.M609600200. [DOI] [PubMed] [Google Scholar]
- 177.Gui J, Liu B, Cao G, Lipchik AM, Perez M, Dekan Z, Mobli M, Daly NL, Alewood PF, Parker LL, King GF, Zhou Y, Jordt SE, Nitabach MN. A Tarantula-Venom Peptide Antagonizes the TRPA1 Nociceptor Ion Channel by Binding to the S1-S4 Gating Domain. Curr Biol. 2014;24:473–83. doi: 10.1016/j.cub.2014.01.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Ibanez-Tallon I, Wen H, Miwa JM, Xing J, Tekinay AB, Ono F, Brehm P, Heintz N. Tethering naturally occurring peptide toxins for cell-autonomous modulation of ion channels and receptors in vivo. Neuron. 2004;43:305–11. doi: 10.1016/j.neuron.2004.07.015. [DOI] [PubMed] [Google Scholar]
- 179.Wu Y, Cao G, Pavlicek B, Luo X, Nitabach MN. Phase coupling of a circadian neuropeptide with rest/activity rhythms detected using a membrane-tethered spider toxin. PLoS Biol. 2008;6:e273. doi: 10.1371/journal.pbio.0060273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 180.Prommer E. Ziconotide: a new option for refractory pain. Drugs Today (Barc) 2006;42:369–78. doi: 10.1358/dot.2006.42.6.973534. [DOI] [PubMed] [Google Scholar]
- 181.Vetter I, Davis JL, Rash LD, Anangi R, Mobli M, Alewood PF, Lewis RJ, King GF. Venomics: a new paradigm for natural products-based drug discovery. Amino Acids. 2011;40:15–28. doi: 10.1007/s00726-010-0516-4. [DOI] [PubMed] [Google Scholar]
- 182.Klint JK, Senff S, Saez NJ, Seshadri R, Lau HY, Bende NS, Undheim EA, Rash LD, Mobli M, King GF. Production of recombinant disulfide-rich venom peptides for structural and functional analysis via expression in the periplasm of E. coli. PLoS One. 2013;8:e63865. doi: 10.1371/journal.pone.0063865. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 183.Schroeder CI, Rash LD, Vila-Farres X, Rosengren KJ, Mobli M, King GF, Alewood PF, Craik DJ, Durek T. Chemical synthesis, 3D structure, and ASIC binding site of the toxin mambalgin-2. Angewandte Chemie. 2014;53:1017–20. doi: 10.1002/anie.201308898. [DOI] [PubMed] [Google Scholar]
- 184.Terstappen GC, Roncarati R, Dunlop J, Peri R. Screening technologies for ion channel drug discovery. Future Med Chem. 2010;2:715–30. doi: 10.4155/fmc.10.180. [DOI] [PubMed] [Google Scholar]
- 185.Wickenden A, Priest B, Erdemli G. Ion channel drug discovery: challenges and future directions. Future Med Chem. 2012;4:661–79. doi: 10.4155/fmc.12.4. [DOI] [PubMed] [Google Scholar]
- 186.Gee KR, Brown KA, Chen WN, Bishop-Stewart J, Gray D, Johnson I. Chemical and physiological characterization of fluo-4 Ca2+-indicator dyes. Cell Calcium. 2000;27:97–106. doi: 10.1054/ceca.1999.0095. [DOI] [PubMed] [Google Scholar]
- 187.Liu Y, Beck EJ, Flores CM. Validation of a patch clamp screening protocol that simultaneously measures compound activity in multiple states of the voltage-gated sodium channel NaV1.2. Assay and Drug Development Technologies. 2011;9:628–34. doi: 10.1089/adt.2011.0375. [DOI] [PubMed] [Google Scholar]
- 188.Cunningham BC, Wells JA. High-resolution epitope mapping of hGH-receptor interactions by alanine-scanning mutagenesis. Science. 1989;244:1081–1085. doi: 10.1126/science.2471267. [DOI] [PubMed] [Google Scholar]
- 189.Revell JD, Lund PE, Linley JE, Metcalfe J, Burmeister N, Sridharan S, Jones C, Jermutus L, Bednarek M. Potency optimization of Huwentoxin-IV on hNaV1.7: A neurotoxin TTX-S sodium-channel antagonist from the venom of the Chinese bird-eating spider Selenocosmia huwena. Peptides. 2013;44:40–46. doi: 10.1016/j.peptides.2013.03.011. [DOI] [PubMed] [Google Scholar]
- 190.Maggio F, King GF. Scanning mutagenesis of a Janus-faced atracotoxin reveals a bipartite surface patch that is essential for neurotoxic function. J Biol Chem. 2002;277:22806–13. doi: 10.1074/jbc.M202297200. [DOI] [PubMed] [Google Scholar]
- 191.Tedford HW, Gilles N, Menez A, Doering CJ, Zamponi GW, King GF. Scanning mutagenesis of ω-atracotoxin-Hv1a reveals a spatially restricted epitope that confers selective activity against insect calcium channels. J Biol Chem. 2004;279:44133–40. doi: 10.1074/jbc.M404006200. [DOI] [PubMed] [Google Scholar]
- 192.Bulaj G. Integrating the discovery pipeline for novel compounds targeting ion channels. Current opinion in chemical biology. 2008;12:441–7. doi: 10.1016/j.cbpa.2008.07.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 193.Salminen O, Drapeau JA, McIntosh JM, Collins AC, Marks MJ, Grady SR. Pharmacology of alpha-conotoxin MII-sensitive subtypes of nicotinic acetylcholine receptors isolated by breeding of null mutant mice. Mol Pharmacol. 2007;71:1563–71. doi: 10.1124/mol.106.031492. [DOI] [PubMed] [Google Scholar]
- 194.Vink S, Alewood PF. Targeting voltage-gated calcium channels: developments in peptide and small-molecule inhibitors for the treatment of neuropathic pain. British journal of pharmacology. 2012;167:970–89. doi: 10.1111/j.1476-5381.2012.02082.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Craik DJ, Adams DJ. Chemical modification of conotoxins to improve stability and activity. ACS Chem Biol. 2007;2:457–68. doi: 10.1021/cb700091j. [DOI] [PubMed] [Google Scholar]
- 196.Castañeda O, Sotolongo V, Amor AM, Stöcklin R, Anderson AJ, Harvey AL, Engström Å, Wernstedt C, Karlsson E. Characterization of a potassium channel toxin from the Caribbean sea anemone Stichodactyla helianthus. Toxicon. 1995;33:603–613. doi: 10.1016/0041-0101(95)00013-c. [DOI] [PubMed] [Google Scholar]
- 197.Beeton C, Wulff H, Standifer NE, Azam P, Mullen KM, Pennington MW, Kolski-Andreaco A, Wei E, Grino A, Counts DR, Wang PH, LeeHealey CJ, B SA, Sankaranarayanan A, Homerick D, Roeck WW, Tehranzadeh J, Stanhope KL, Zimin P, Havel PJ, Griffey S, Knaus HG, Nepom GT, Gutman GA, Calabresi PA, Chandy KG. Kv1.3 channels are a therapeutic target for T cell-mediated autoimmune diseases. Proc Natl Acad Sci U S A. 2006;103:17414–9. doi: 10.1073/pnas.0605136103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Chi V, Pennington MW, Norton RS, Tarcha EJ, Londono LM, Sims-Fahey B, Upadhyay SK, Lakey JT, Iadonato S, Wulff H, Beeton C, Chandy KG. Development of a sea anemone toxin as an immunomodulator for therapy of autoimmune diseases. Toxicon. 2012;59:529–46. doi: 10.1016/j.toxicon.2011.07.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Galea CA, Nguyen HM, Chandy KG, Smith BJ, Norton RS. Domain structure and function of matrix metalloprotease 23 (MMP23): role in potassium channel trafficking. Cell Mol Life Sci. 2014;71:1191–210. doi: 10.1007/s00018-013-1431-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Nguyen HM, Galea CA, Schmunk G, Smith BJ, Edwards RA, Norton RS, Chandy KG. Intracellular trafficking of the KV1.3 potassium channel is regulated by the prodomain of a matrix metalloprotease. J Biol Chem. 2013;288:6451–64. doi: 10.1074/jbc.M112.421495. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Quik M, McIntosh JM. Striatal alpha6* nicotinic acetylcholine receptors: potential targets for Parkinson's disease therapy. J Pharmacol Exp Ther. 2006;316:481–9. doi: 10.1124/jpet.105.094375. [DOI] [PubMed] [Google Scholar]
- 202.Olivera BM, Quik M, Vincler M, McIntosh JM. Subtype-selective conopeptides targeted to nicotinic receptors: Concerted discovery and biomedical applications. Channels (Austin) 2008;2 doi: 10.4161/chan.2.2.6276. [DOI] [PubMed] [Google Scholar]
- 203.Azam L, McIntosh JM. Alpha-conotoxins as pharmacological probes of nicotinic acetylcholine receptors. Acta Pharmacol Sin. 2009;30:771–83. doi: 10.1038/aps.2009.47. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 204.Bordia T, Parameswaran N, Fan H, Langston JW, McIntosh JM, Quik M. Partial recovery of striatal nicotinic receptors in 1-methyl-4-phenyl-1,2,3,6-tetrahydropyridine (MPTP)-lesioned monkeys with chronic oral nicotine. J Pharmacol Exp Ther. 2006;319:285–92. doi: 10.1124/jpet.106.106997. [DOI] [PubMed] [Google Scholar]
- 205.McIntosh JM, Azam L, Staheli S, Dowell C, Lindstrom JM, Kuryatov A, Garrett JE, Marks MJ, Whiteaker P. Analogs of a-conotoxin MII are selective for a6-containing nicotinic acetylcholine receptors. Mol Pharmacol. 2004;65:944–952. doi: 10.1124/mol.65.4.944. [DOI] [PubMed] [Google Scholar]
- 206.Quik M, Bordia T, Forno L, McIntosh JM. Loss of a-conotoxin MII- and A85380-sensitive nicotinic receptors in Parkinson's disease striatum. J Neurochem. 2004;88:668–79. doi: 10.1111/j.1471-4159.2004.02177.x. [DOI] [PubMed] [Google Scholar]
- 207.Martin MF, Rochat H, Marchot P, Bougis PE. Use of high performance liquid chromatography to demonstrate quantitative variation in components of venom from the scorpion Androctonus australis Hector. Toxicon. 1987;25:569–73. doi: 10.1016/0041-0101(87)90293-5. [DOI] [PubMed] [Google Scholar]
- 208.Martin MF, Garcia y Perez LG, el Ayeb M, Kopeyan C, Bechis G, Jover E, Rochat H. Purification and chemical and biological characterizations of seven toxins from the Mexican scorpion, Centruroides suffusus suffusus. J Biol Chem. 1987;262:4452–9. [PubMed] [Google Scholar]
- 209.Cartier GE, Yoshikami D, Gray WR, Luo S, Olivera BM, McIntosh JM. A new a-conotoxin which targets a3b2 nicotinic acetylcholine receptors. J Biol Chem. 1996;271:7522–7528. doi: 10.1074/jbc.271.13.7522. [DOI] [PubMed] [Google Scholar]
- 210.Wang J, Salata JJ, Bennett PB. Saxitoxin is a gating modifier of HERG K+ channels. J. Gen. Physiol. 2003;121:583–598. doi: 10.1085/jgp.200308812. [DOI] [PMC free article] [PubMed] [Google Scholar]




