426
Views
1
CrossRef citations to date
0
Altmetric
Review

Spider Venomics: Implications for Drug Discovery

, , , &
Pages 1699-1714 | Published online: 18 Nov 2014

References

  • Fry BG , RoelantsK, ChampagneDEet al. The toxicogenomic multiverse: convergent recruitment of proteins into animal venoms. Annu. Rev. Genomics Hum. Genet.10 (1), 483–511 (2009).
  • von Reumont BM , BlankeA, RichterS, AlvarezF, BleidornC, JennerRA. The first venomous crustacean revealed by transcriptomics and functional morphology: remipede venom glands express a unique toxin cocktail dominated by enzymes and a neurotoxin. Mol. Biol. Evol.31 (1), 48–58 (2014).
  • Berkov A , RodriguezN, CentenoP. Convergent evolution in the antennae of a cerambycid beetle, Onychocerus albitarsis, and the sting of a scorpion. Naturwissenschaften95 (3), 257–261 (2008).
  • Platnick NI . Advances in spider taxonomy, 1992–1995: with redescriptions 1940–1980. New York Entomological Society & The American Museum of Natural History, NY, USA (2011). Updated information and statistics available athttp://www.wsc.nmbe.ch/statistics/
  • Escoubas P , BosmansF. Spider peptide toxins as leads for drug development. Expert Opin. Drug Discov.2 (6), 823–835 (2007).
  • King GF , HardyMC. Spider-venom peptides: structure, pharmacology, and potential for control of insect pests. Annu. Rev. Entomol.58, 475–496 (2013).
  • Rash LD , HodgsonWC. Pharmacology and biochemistry of spider venoms. Toxicon40 (3), 225–254 (2002).
  • King GF . The wonderful world of spiders: preface to the special Toxicon issue on spider venoms. Toxicon43 (5), 471–475 (2004).
  • Escoubas P , SollodB, KingGF. Venom landscapes: mining the complexity of spider venoms via a combined cDNA and mass spectrometric approach. Toxicon47 (6), 650–663 (2006).
  • Kuhn-Nentwig L , StöcklinR, NentwigW. Venom composition and strategies in spiders: is everything possible?Adv. Insect Physiol.40, 1–86 (2011).
  • King GF . Venoms to drugs: translating venom peptides into therapeutics. Aust. Biochem.44 (3), 13–15 (2013).
  • King GF . Venoms as a platform for human drugs: translating toxins into therapeutics. Expert Opin. Biol. Ther.11 (11), 1469–1484 (2011).
  • Vetter I , DavisJL, RashLDet al. Venomics: a new paradigm for natural products-based drug discovery. Amino Acids40, 15–28 (2011).
  • Escoubas P , QuintonL, NicholsonGM. Venomics: unravelling the complexity of animal venoms with mass spectrometry. J. Mass Spectrom.43 (3), 279–295 (2008).
  • Dutertre S , JinAH, KaasQ, JonesA, AlewoodPF, LewisRJ. Deep venomics reveals the mechanism for expanded peptide diversity in cone snail venom. Mol. Cell. Proteomics12 (2), 312–329 (2013).
  • Herzig V , WoodDLA, NewellFet al. ArachnoServer 2.0, an updated online resource for spider toxin sequences and structures Nucleic Acids Res. 39, D653–D657 (2011).
  • Pallaghy PK , NielsenKJ, CraikDJ, NortonRS. A common structural motif incorporating a cystine knot and a triple-stranded β-sheet in toxic and inhibitory polypeptides. Protein Sci.3, 1833–1839 (1994).
  • King GF , TedfordHW, MaggioF. Structure and function of insecticidal neurotoxins from Australian funnel-web spiders. J. Toxicol. Toxin Rev.21, 359–389 (2002).
  • Vassilevski AA , KozlovSA, GrishinEV. Molecular diversity of spider venom. Biochemistry (Moscow)74 (13), 1505–1534 (2009).
  • Saez NJ , SenffS, JensenJEet al. Spider-venom peptides as therapeutics. Toxins2, 2851–2871 (2010).
  • Colgrave ML , CraikDJ. Thermal, chemical, and enzymatic stability of the cyclotide kalata B1: the importance of the cyclic cystine knot. Biochemistry43 (20), 5965–5975 (2004).
  • Fletcher JI , SmithR, O’DonoghueSIet al. The structure of a novel insecticidal neurotoxin, ω-atracotoxin-HV1, from the venom of an Australian funnel web spider. Nat. Struct. Biol.4, 559–566 (1997).
  • Perola E . An analysis of the binding efficiencies of drugs and their leads in successful drug discovery programs. J. Med. Chem.53 (7), 2986–2997 (2010).
  • Gui J , LiuB, CaoGet al. A tarantula-venom peptide antagonizes the TRPA1 nociceptor ion channel by binding to the S1–S4 gating domain. Curr. Biol.24 (5), 473–483 (2014).
  • Frontali N , CeccarelliB, GorioAet al. Purification from black widow spider venom of a protein factor causing the depletion of synaptic vesicles at neuromuscular junctions. J. Cell Biol.68 (3), 462–479 (1976).
  • Sheumack DD , ClaassensR, WhiteleyNM, HowdenME. Complete amino acid sequence of a new type of lethal neurotoxin from the venom of the funnel-web spider Atrax robustus. FEBS Lett.181 (1), 154–156 (1985).
  • Brown MR , SheumackDD, TylerMI, HowdenME. Amino acid sequence of versutoxin, a lethal neurotoxin from the venom of the funnel-web spider Atrax versutus. Biochem. J.250 (2), 401–405 (1988).
  • Kiyatkin NI , DulubovaIE, ChekhovskayaIA, GrishinEV. Cloning and structure of cDNA encoding α-latrotoxin from black widow spider venom. FEBS Lett.270 (1–2), 127–131 (1990).
  • Frohman MA , DushMK, MartinGR. Rapid production of full-length cDNAs from rare transcripts: amplification using a single gene-specific oligonucleotide primer. Proc. Natl Acad. Sci. USA85 (23), 8998–9002 (1988).
  • Wang X-H , ConnorM, WilsonDet al. Discovery and structure of a potent and highly specific blocker of insect calcium channels. J. Biol. Chem.276 (43), 40306–40312 (2001).
  • Diao JB , LinY, TangJZ, LiangSP. cDNA sequence analysis of seven peptide toxins from the spider Selenocosmia huwena. Toxicon42 (7), 715–723 (2003).
  • Sollod BL , WilsonD, ZhaxybayevaO, GogartenJP, DrinkwaterR, KingGF. Were arachnids the first to use combinatorial peptide libraries?Peptides26 (1), 131–139 (2005).
  • Kozlov S , GrishinEV. Classification of spider neurotoxins using structural motifs by primary structure features. Single residue distribution analysis and pattern analysis techniques. Toxicon46 (6), 672–686 (2005).
  • Chen J , ZhaoL, JiangLet al. Transcriptome analysis revealed novel possible venom components and cellular processes of the tarantula Chilobrachys jingzhao venom gland. Toxicon52 (7), 794–806 (2008).
  • Tang X , ZhangY, HuWet al. Molecular diversification of peptide toxins from the tarantula Haplopelma hainanum (Ornithoctonus hainana) venom based on transcriptomic, peptidomic, and genomic analyses. J. Proteome Res.9 (5), 2550–2564 (2010).
  • Jiang L , ChenJ, PengL, ZhangY, XiongX, LiangS. Genomic organization and cloning of novel genes encoding toxin-like peptides of three superfamilies from the spider Orinithoctonus huwena. Peptides29 (10), 1679–1684 (2008).
  • Gremski LH , da SilveiraRB, ChaimOMet al. A novel expression profile of the Loxosceles intermedia spider venomous gland revealed by transcriptome analysis. Mol. Biosyst.6 (12), 2403–2416 (2010).
  • Fernandes-Pedrosa MdF , Junqueira-de-AzevedoIdLM, Goncalves-de-AndradeRMet al. Transcriptome analysis of Loxosceles laeta (Araneae, Sicariidae) spider venomous gland using expressed sequence tags. BMC Genomics9, 279 (2008).
  • Zhang Y , ChenJ, TangXet al. Transcriptome analysis of the venom glands of the Chinese wolf spider Lycosa singoriensis. Zoology113 (1), 10–18 (2010).
  • Diego-Garcia E , PeigneurS, WaelkensE, DebaveyeS, TytgatJ. Venom components from Citharischius crawshayi spider (Family Theraphosidae): exploring transcriptome, venomics, and function. Cell. Mol. Life Sci.67 (16), 2799–2813 (2010).
  • Wong ES , HardyMC, WoodD, BaileyT, KingGF. SVM-based prediction of propeptide cleavage sites in spider toxins identifies toxin innovation in an Australian tarantula. PLoS ONE8 (7), e66279 (2013).
  • Undheim EA , SunagarK, HerzigVet al. A proteomics and transcriptomics investigation of the venom from the barychelid spider Trittame loki (brush-foot trapdoor). Toxins5 (12), 2488–2503 (2013).
  • He Q , DuanZ, YuY, LiuZ, LiuZ, LiangS. The venom gland transcriptome of Latrodectus tredecimguttatus revealed by deep sequencing and cDNA library analysis. PLoS ONE8 (11), e81357 (2013).
  • Sanggaard KW , BechsgaardJS, FangXet al. Spider genomes provide insight into composition and evolution of venom and silk. Nat. Commun.5, 3765 (2014).
  • Gregory TR , ShorthouseDP. Genome sizes of spiders. J. Heredity94, 285–290 (2003).
  • Davis J , JonesA, LewisRJ. Remarkable inter- and intra-species complexity of conotoxins revealed by LC/MS. Peptides30 (7), 1222–1227 (2009).
  • Bandeira N , ClauserKR, PevznerPA. Shotgun protein sequencing: assembly of peptide tandem mass spectra from mixtures of modified proteins. Mol. Cell. Proteomics6 (7), 1123–1134 (2007).
  • Ueberheide BM , FenyoD, AlewoodPF, ChaitBT. Rapid sensitive analysis of cysteine rich peptide venom components. Proc. Natl Acad. Sci. USA106 (17), 6910–6915 (2009).
  • Perkins DN , PappinDJC, CreasyDM, CottrellJS. Probability-based protein identification by searching sequence databases using mass spectrometry data. Electrophoresis20 (18), 3551–3567 (1999).
  • Craig R , BeavisRC. TANDEM: matching proteins with tandem mass spectra. Bioinformatics20 (9), 1466–1467 (2004).
  • Zobel-Thropp PA , CorreaSM, GarbJE, BinfordGJ. Spit and venom from Scytodes spiders: a diverse and distinct cocktail. J. Proteome Res.13 (2), 817–835 (2013).
  • Duan Z , CaoR, JiangL, LiangS. A combined de novo protein sequencing and cDNA library approach to the venomic analysis of Chinese spider Araneus ventricosus. J. Proteomics78, 416–427 (2013).
  • Craik DJ , AdamsDJ. Chemical modification of conotoxins to improve stability and activity. ACS Chem. Biol.2 (7), 457–468 (2007).
  • Muttenthaler M , AkondiKB, AlewoodPF. Structure-activity studies on α-conotoxins. Curr. Pharm. Des.17 (38), 4226–4241 (2011).
  • Revell JD , LundPE, LinleyJEet al. 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. Peptides44, 40–46 (2013).
  • Dawson PE , MuirTW, Clark-LewisI, KentSBH. Synthesis by native chemical ligation. Science266, 776–779 (1994).
  • Jensen JE , DurekT, AlewoodPF, AdamsDJ, KingGF, RashLD. Chemical synthesis and folding of APETx2, a potent and selective inhibitor of acid sensing ion channel 3. Toxicon54 (1), 56–61 (2009).
  • Durek T , VetterI, WangC-IAet al. Chemical engineering and structural and pharmacological characterization of the α-scorpion toxin OD1. ACS Chem. Biol.8 (6), 1215–1222 (2013).
  • Schroeder CI , RashLD, Vila-FarrésXet al. Chemical synthesis, 3D structure, and ASIC binding site of the toxin mambalgin-2. Angew. Chem. Int. Ed.53 (4), 1017–1020 (2014).
  • Bulaj G , OliveraBM. Folding of conotoxins: formation of the native disulfide bridges during chemical synthesis and biosynthesis of Conus peptides. Antioxid. Redox Signal.10 (1), 141–155 (2008).
  • Anangi R , ChenC-Y, ChengC-Het al. Expression of snake venom toxins in Pichia pastoris. Toxin Rev.26, 169–187 (2007).
  • Makrides SC . Strategies for achieving high-level expression of genes in Escherichia coli. Microbiol. Rev.60 (3), 512–538 (1996).
  • Baneyx F , MujacicM. Recombinant protein folding and misfolding in Escherichia coli. Nat. Biotech.22 (11), 1399–1408 (2004).
  • Tedford HW , FletcherJI, KingGF. Functional significance of the β-hairpin in the insecticidal neurotoxin ω-atracotoxin-Hv1a. J. Biol. Chem.276, 26568–26576 (2001).
  • Klint JK , SenffS, SaezNJet al. Production of recombinant disulfide-rich venom peptides for structural and functional analysis via expression in the periplasm of E. coli. PLoS ONE8 (5), e63865 (2013).
  • Ito K , InabaK. The disulfide bond formation (Dsb) system. Curr. Opin. Struct. Biol.18 (4), 450–458 (2008).
  • Heras B , ShouldiceSR, TotsikaM, ScanlonMJ, SchembriMA, MartinJL. DSB proteins and bacterial pathogenicity. Nat. Rev. Microbiol.7 (3), 215–225 (2009).
  • Kwan AH , MobliM, GooleyPR, KingGF, MackayJP. Macromolecular NMR spectroscopy for the non-spectroscopist. FEBS J.278 (5), 687–703 (2011).
  • Dawson RJ , BenzJ, StohlerPet al. Structure of the acid-sensing ion channel 1 in complex with the gating modifier psalmotoxin 1. Nat. Commun.3, 936 (2012).
  • Baconguis I , GouauxE. Structural plasticity and dynamic selectivity of acid-sensing ion channel-spider toxin complexes. Nature489 (7416), 400–405 (2012).
  • Cao E , LiaoM, ChengY, JuliusD. TRPV1 structures in distinct conformations reveal activation mechanisms. Nature504 (7478), 113–118 (2013).
  • Mobli M , MaciejewskiMW, GrykMR, HochJC. An automated tool for maximum entropy reconstruction of biomolecular NMR spectra. Nat. Methods4 (6), 467–468 (2007).
  • Mobli M , SternAS, BermelW, KingGF, HochJC. A non-uniformly sampled 4D HCC(CO)NH-TOCSY experiment processed using maximum entropy for rapid protein sidechain assignment. J. Magn. Reson.204, 160–164 (2010).
  • Güntert P . Automated NMR structure calculation with CYANA. Methods Mol. Biol.278, 353–378 (2004).
  • Saez NJ , MobliM, BieriMet al. A dynamic pharmacophore drives the interaction between psalmotoxin-1 and the putative drug target acid-sensing ion channel 1a. Mol. Pharmacol.80, 796–808 (2011).
  • Bende NS , KangE, HerzigVet al. The insecticidal neurotoxin Aps III is an atypical knottin peptide that potently blocks insect voltage-gated sodium channels. Biochem. Pharmacol.85 (10), 1542–1554 (2013).
  • Cunningham BC , WellsJA. High-resolution epitope mapping of hGH-receptor interactions by alanine-scanning mutagenesis. Science244 (4908), 1081–1085 (1989).
  • Maggio F , KingGF. Scanning mutagenesis of a Janus-faced atracotoxin reveals a bipartite surface patch that is essential for neurotoxic function. J. Biol. Chem.277, 22806–22813 (2002).
  • Tedford HW , GillesN, MénezA, DoeringCJ, ZamponiGW, KingGF. Scanning mutagenesis of ω-atracotoxin-Hv1a reveals a spatially restricted epitope that confers selective activity against invertebrate calcium channels. J. Biol. Chem.279, 44133–44140 (2004).
  • Corzo G , SaboJK, BosmansFet al. Solution structure and alanine scan of a spider toxin that affects the activation of mammalian voltage-gated sodium channels. J. Biol. Chem.282 (7), 4643–4652 (2007).
  • Estrada G , VillegasE, CorzoG. Spider venoms: a rich source of acylpolyamines and peptides as new leads for CNS drugs. Nat. Prod. Rep.24 (1), 145–161 (2007).
  • Windley MJ , HerzigV, DziemborowiczSA, HardyMC, KingGF, NicholsonGM. Spider-venom peptides as bioinsecticides. Toxins4 (3), 191–227 (2012).
  • Catterall WA , GoldinAL, WaxmanSG. International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol. Rev.57 (4), 397–409 (2005).
  • Yu FH , CatterallWA. Overview of the voltage-gated sodium channel family. Genome Biol.4, 207 (2003).
  • King GF , EscoubasP, NicholsonGM. Peptide toxins that selectively target insect NaV and CaV channels. Channels2 (2), 100–116 (2008).
  • Waxman SG , Dib-HajjS. Erythermalgia: molecular basis for an inherited pain syndrome. Trends Mol. Med.11 (12), 555–562 (2005).
  • Rush AM , Dib-HajjSD, LiuSJ, CumminsTR, BlackJA, WaxmanSG. A single sodium channel mutation produces hyperor hypoexcitability in different types of neurons. Proc. Natl Acad. Sci. USA103, 8245–8250 (2006).
  • Jarecki BW , SheetsPL, JacksonJO, CumminsTR. Paroxysmal extreme pain disorder mutations within the D3/S4-S5 linker of Nav1.7 cause moderate destabilization of fast inactivation. J. Physiol.586 (17), 4137–4153 (2008).
  • Cox JJ , ReimannF, NicholasAKet al. An SCN9A channelopathy causes congenital inability to experience pain. Nature444 (7121), 894–898 (2006).
  • Weiss J , PyrskiM, JacobiEet al. Loss-of-function mutations in sodium channel Nav1.7 cause anosmia. Nature472 (7342), 186–190 (2011).
  • Rupasinghe DB , KnappO, BlomsterLVet al. Localization of NaV1.7 in the normal and injured rodent olfactory system indicates a critical role in olfaction, pheromone sensing and immune function. Channels6, 103–110 (2012).
  • Reimann F , CoxJJ, BelferIet al. Pain perception is altered by a nucleotide polymorphism in SCN9A. Proc. Natl Acad. Sci. USA107 (11), 5148–5153 (2010).
  • Duan G , XiangG, ZhangX, YuanR, ZhanH, QiD. A single-nucleotide polymorphism in SCN9A may decrease postoperative pain sensitivity in the general population. Anesthesiology118, 436–442 (2013).
  • Clare JJ . Targeting voltage-gated sodium channels for pain therapy. Expert Opin. Investig. Drugs19 (1), 45–62 (2010).
  • England S , RawsonD. Isoform-selective voltage-gated Na+ channel modulators as next-generation analgesics. Future Med. Chem.2 (5), 775–790 (2010).
  • Catterall WA . From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels. Neuron26 (1), 13–25 (2000).
  • Krafte DS , BannonAW. Sodium channels and nociception: recent concepts and therapeutic opportunities. Curr. Opin. Pharmacol.8 (1), 50–56 (2008).
  • Klint JK , SenffS, RupasingheDBet al. Spider-venom peptides that target voltage-gated sodium channels: pharmacological tools and potential therapeutic leads. Toxicon60 (4), 478–491 (2012).
  • Schmalhofer WA , CalhounJ, BurrowsRet al. ProTx-II, a selective inhibitor of NaV1.7 sodium channels, blocks action potential propagation in nociceptors. Mol. Pharmacol.74 (5), 1476–1484 (2008).
  • Bosmans F , RashL, ZhuSYet al. Four novel tarantula toxins as selective modulators of voltage-gated sodium channel subtypes. Mol. Pharmacol.69 (2), 419–429 (2006).
  • Ono S , KimuraT, KuboT. Characterization of voltage-dependent calcium channel blocking peptides from the venom of the tarantula Grammostola rosea. Toxicon58 (3), 265–276 (2011).
  • King GF , GentzMC, EscoubasP, NicholsonGM. A rational nomenclature for naming peptide toxins from spiders and other venomous animals. Toxicon52 (2), 264–276 (2008).
  • ArachnoServer . www.arachnoserver.org/toxincard.html?id=AS-Card-Number
  • King GF . Modulation of insect CaV channels by peptidic spider toxins. Toxicon49 (4), 513–530 (2007).
  • Altier C , ZamponiGW. Targeting Ca2+ channels to treat pain: T-type versus N-type. Trends Pharmacol. Sci.25 (9), 465–470 (2004).
  • McGivern JG , McDonoughSI. Voltage-gated calcium channels as targets for the treatment of chronic pain. Curr. Drug Targets CNS Neurol. Disord.3 (6), 457–478 (2004).
  • Deng M , LuoX, XiaoYet al. Huwentoxin-XVI, an analgesic, highly reversible mammalian N-type calcium channel antagonist from Chinese tarantula Ornithoctonus huwena. Neuropharmacology79, 657–667 (2014).
  • Miljanich GP . Ziconotide: neuronal calcium channel blocker for treating severe chronic pain. Curr. Med. Chem.11 (23), 3029–3040 (2004).
  • McNamara CR , Mandel-BrehmJ, BautistaDMet al. TRPA1 mediates formalin-induced pain. Proc. Natl Acad. Sci. USA104 (33), 13525–13530 (2007).
  • Lewis RJ , DutertreS, VetterI, ChristieMJ. Conus venom peptide pharmacology. Pharmacol. Rev.64 (2), 259–298 (2012).
  • Kellenberger S , SchildL. Epithelial sodium channel/degenerin family of ion channels: a variety of functions for a shared structure. Physiol. Rev.82 (3), 735–767 (2002).
  • Lingueglia E . Acid-sensing ion channels in sensory perception. J. Biol. Chem.282 (24), 17325–17329 (2007).
  • Sluka KA , WinterOC, WemmieJA. Acid-sensing ion channels: a new target for pain and CNS diseases. Curr. Opin. Drug Discov. Dev.12 (5), 693–704 (2009).
  • Voilley N . Acid-sensing ion channels (ASICs): new targets for the analgesic effects of non-steroid anti-inflammatory drugs (NSAIDs). Curr. Drug Targets Inflamm. Allergy3 (1), 71–79 (2004).
  • Gwanyanya A , MacianskieneR, MubagwaK. Insights into the effects of diclofenac and other non-steroidal anti-inflammatory agents on ion channels. J. Pharm. Pharmacol.64 (10), 1359–1375 (2012).
  • Escoubas P , De WeilleJR, LecoqAet al. Isolation of a tarantula toxin specific for a class of proton-gated Na+ channels. J. Biol. Chem.275 (33), 25116–25121 (2000).
  • Mazzuca M , HeurteauxC, AllouiAet al. A tarantula peptide against pain via ASIC1a channels and opioid mechanisms. Nat. Neurosci.10 (8), 943–945 (2007).
  • Smith HS . Intrathecal drug delivery. Pain Physician11, S89–S104 (2008).
  • Tedford HW , SteinbaughBA, BaoLet al. In silico screening for compounds that match the pharmacophore of ω-hexatoxin-Hv1a leads to discovery and optimization of a novel class of insecticides. Pesticide Biochem. Physiol.106, 124–140 (2013).
  • Moskowitz MA , LoEH, IadecolaC. The science of stroke: mechanisms in search of treatments. Neuron67 (2), 181–198 (2010).
  • Woodruff TM , ThundyilJ, TangSC, SobeyCG, TaylorSM, ArumugamTV. Pathophysiology, treatment, and animal and cellular models of human ischemic stroke. Mol. Neurodegener.6 (1), 11 (2011).
  • Liu R , YuanH, YuanF, YangSH. Neuroprotection targeting ischemic penumbra and beyond for the treatment of ischemic stroke. Neurol. Res.34 (4), 331–337 (2012).
  • Xiong ZG , ChuXP, SimonRP. Acid sensing ion channels – novel therapeutic targets for ischemic brain injury. Front. Biosci.12, 1376–1386 (2007).
  • Xiong ZG , ZhuXM, ChuXPet al. Neuroprotection in ischemia: blocking calcium-permeable acid-sensing ion channels. Cell118 (6), 687–698 (2004).
  • Isaev NK , StelmashookEV, PlotnikovEYet al. Role of acidosis, NMDA receptors, and acid-sensitive ion channel 1a (ASIC1a) in neuronal death induced by ischemia. Biochemistry (Moscow)73 (11), 1171–1175 (2008).
  • Leng TD , XiongZG. The pharmacology and therapeutic potential of small molecule inhibitors of acid-sensing ion channels in stroke intervention. Acta Pharmacol. Sin.34 (1), 33–38 (2013).
  • Pignataro G , SimonRP, XiongZG. Prolonged activation of ASIC1a and the time window for neuroprotection in cerebral ischaemia. Brain130, 151–158 (2007).
  • Ikonomidou C , TurskiL. Why did NMDA receptor antagonists fail clinical trials for stroke and traumatic brain injury?Lancet Neurol.1 (6), 383–386 (2002).
  • Skaper SD , DebettoP, GiustiP. The P2X7 purinergic receptor: from physiology to neurological disorders. FASEB J.24 (2), 337–345 (2010).
  • Nicke A , BaumertHG, RettingerJet al. P2X1 and P2X3 receptors form stable trimers: a novel structural motif of ligand-gated ion channels. EMBO J.17 (11), 3016–3028 (1998).
  • Jarvis MF , BurgardEC, McGaraughtySet al. A-317491, a novel potent and selective nonnucleotide antagonist of P2X3 and P2X2/3 receptors, reduces chronic inflammatory and neuropathic pain in the rat. Proc. Natl Acad. Sci. USA99 (26), 17179–17184 (2002).
  • Hausmann R , RettingerJ, GerevichZet al. The suramin analog 4,4 ’,4 ‘’,4 ‘‘‘-(carbonylbis(imino-5,1,3-benzenetriylbis (carbonylimino)))tetra-kis-benzenesulfonic acid (NF110) potently blocks P2X3 receptors: subtype selectivity is determined by location of sulfonic acid groups. Mol. Pharmacol.69 (6), 2058–2067 (2006).
  • Jung KY , MoonHD, LeeGE, LimHH, ParkCS, KimYC. Structure-activity relationship studies of spinorphin as a potent and selective human P2X3 receptor antagonist. J. Med. Chem.50 (18), 4543–4547 (2007).
  • Grishin EV , SavchenkoGA, VassilevskiAAet al. Novel peptide from spider venom inhibits P2X3 receptors and inflammatory pain. Ann. Neurol.67 (5), 680–683 (2010).
  • Cohen E , QuistadGB. Cytotoxic effects of arthropod venoms on various cultured cells. Toxicon36 (2), 353–358 (1998).
  • Gao L , ShanBE, ChenJ, LiuJH, SongDX, ZhuBC. Effects of spider Macrothele raven venom on cell proliferation and cytotoxicity in HeLa cells. Acta Pharmacol. Sin.26 (3), 369–376 (2005).
  • Gao L , YuS, WuY, ShanB. Effect of spider venom on cell apoptosis and necrosis rates in MCF-7 cells. DNA Cell Biol.26, 485–489 (2007).
  • Liu Z , ZhaoY, LiJet al. The venom of the spider Macrothele raveni induces apoptosis in the myelogenous leukemia K562 cell line. Leuk. Res.36 (8), 1063–1066 (2012).
  • Bubien JK , JiHL, GillespieGYet al. Cation selectivity and inhibition of malignant glioma Na+ channels by Psalmotoxin 1. Am. J. Physiol. Cell Physiol.287, C1282–C1291 (2004).
  • Li M , InoueK, BraniganDet al. Acid-sensing ion channels in acidosis-induced injury of human brain neurons. J. Cereb. Blood Flow Metab.30 (6), 1247–1260 (2010).
  • Deshane J , GarnerCC, SontheimerH. Chlorotoxin inhibits glioma cell invasion via matrix metalloproteinase-2. J. Biol. Chem.278 (6), 4135–4144 (2003).
  • Kesavan K , RatliffJ, JohnsonEWet al. Annexin A2 is a molecular target for TM601, a peptide with tumor-targeting and anti-angiogenic effects. J. Biol. Chem.285 (7), 4366–4374 (2010).
  • Veiseh M , GabikianP, BahramiSBet al. Tumor paint: a chlorotoxin: Cy5.5 bioconjugate for intraoperative visualization of cancer foci. Cancer Res.67 (14), 6882–6888 (2007).
  • Stroud MR , HansenSJ, OlsonJM. In vivo bio-imaging using chlorotoxin-based conjugates. Curr. Pharm. Des.17 (38), 4362–4371 (2011).

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.