Skip to main content

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • Article
  • Published:

Selective spider toxins reveal a role for the Nav1.1 channel in mechanical pain

Abstract

Voltage-gated sodium (Nav) channels initiate action potentials in most neurons, including primary afferent nerve fibres of the pain pathway. Local anaesthetics block pain through non-specific actions at all Nav channels, but the discovery of selective modulators would facilitate the analysis of individual subtypes of these channels and their contributions to chemical, mechanical, or thermal pain. Here we identify and characterize spider (Heteroscodra maculata) toxins that selectively activate the Nav1.1 subtype, the role of which in nociception and pain has not been elucidated. We use these probes to show that Nav1.1-expressing fibres are modality-specific nociceptors: their activation elicits robust pain behaviours without neurogenic inflammation and produces profound hypersensitivity to mechanical, but not thermal, stimuli. In the gut, high-threshold mechanosensitive fibres also express Nav1.1 and show enhanced toxin sensitivity in a mouse model of irritable bowel syndrome. Together, these findings establish an unexpected role for Nav1.1 channels in regulating the excitability of sensory nerve fibres that mediate mechanical pain.

This is a preview of subscription content, access via your institution

Access options

Buy this article

Prices may be subject to local taxes which are calculated during checkout

Figure 1: Hm1a selectively targets Nav1.1 in sensory neurons.
Figure 2: Hm1a targets S3b–S4 and S1–S2 loops in DIV to inhibit fast inactivation.
Figure 3: Nav1.1 is expressed by myelinated, non-C-fibre neurons in sensory ganglia.
Figure 4: Hm1a elicits non-inflammatory pain and bilateral mechanical allodynia.
Figure 5: Colonic afferents display increased sensitivity to Hm1a in a model of IBS.

Similar content being viewed by others

Accession codes

Data deposits

Sequences and activity profiles for Hm1a and Hm1b can be found in the ArachnoServer database49 under accession numbers AS000224 and AS002363, respectively.

References

  1. Basbaum, A. I., Bautista, D. M., Scherrer, G. & Julius, D. Cellular and molecular mechanisms of pain. Cell 139, 267–284 (2009)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  2. Dib-Hajj, S. D., Yang, Y., Black, J. A. & Waxman, S. G. The Nav1.7 sodium channel: from molecule to man. Nat. Rev. Neurosci. 14, 49–62 (2013)

    Article  CAS  PubMed  Google Scholar 

  3. Faber, C. G. et al. Gain-of-function Nav1.8 mutations in painful neuropathy. Proc. Natl Acad. Sci. USA 109, 19444–19449 (2012)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  4. Fertleman, C. R. et al. SCN9A mutations in paroxysmal extreme pain disorder: allelic variants underlie distinct channel defects and phenotypes. Neuron 52, 767–774 (2006)

    Article  CAS  PubMed  Google Scholar 

  5. Hoeijmakers, J. G., Faber, C. G., Merkies, I. S. & Waxman, S. G. Painful peripheral neuropathy and sodium channel mutations. Neurosci. Lett. 596, 51–59 (2015)

    Article  CAS  PubMed  Google Scholar 

  6. Zhang, X. Y. et al. Gain-of-function mutations in SCN11A cause familial episodic pain. Am. J. Hum. Genet. 93, 957–966 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  7. Fukuoka, T. et al. Comparative study of the distribution of the α-subunits of voltage-gated sodium channels in normal and axotomized rat dorsal root ganglion neurons. J. Comp. Neurol. 510, 188–206 (2008)

    Article  CAS  PubMed  Google Scholar 

  8. Black, J. A., Liu, S., Tanaka, M., Cummins, T. R. & Waxman, S. G. Changes in the expression of tetrodotoxin-sensitive sodium channels within dorsal root ganglia neurons in inflammatory pain. Pain 108, 237–247 (2004)

    Article  CAS  PubMed  Google Scholar 

  9. Fukuoka, T., Miyoshi, K. & Noguchi, K. De novo expression of Nav1.7 in injured putative proprioceptive afferents: Multiple tetrodotoxin-sensitive sodium channels are retained in the rat dorsal root after spinal nerve ligation. Neuroscience 284, 693–706 (2015)

    Article  CAS  PubMed  Google Scholar 

  10. Dib-Hajj, S. D., Cummins, T. R., Black, J. A. & Waxman, S. G. Sodium channels in normal and pathological pain. Annu. Rev. Neurosci. 33, 325–347 (2010)

    Article  CAS  PubMed  Google Scholar 

  11. Wang, W., Gu, J., Li, Y. Q. & Tao, Y. X. Are voltage-gated sodium channels on the dorsal root ganglion involved in the development of neuropathic pain? Mol. Pain 7, 16 (2011)

    PubMed  PubMed Central  Google Scholar 

  12. Catterall, W. A., Kalume, F. & Oakley, J. C. Nav1.1 channels and epilepsy. J. Physiol. (Lond.) 588, 1849–1859 (2010)

    Article  CAS  Google Scholar 

  13. Cheah, C. S. et al. Specific deletion of Nav1.1 sodium channels in inhibitory interneurons causes seizures and premature death in a mouse model of Dravet syndrome. Proc. Natl Acad. Sci. USA 109, 14646–14651 (2012)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  14. Han, S. et al. Autistic-like behaviour in Scn1a+/− mice and rescue by enhanced GABA-mediated neurotransmission. Nature 489, 385–390 (2012)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  15. Verret, L. et al. Inhibitory interneuron deficit links altered network activity and cognitive dysfunction in Alzheimer model. Cell 149, 708–721 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  16. Gargus, J. J. & Tournay, A. Novel mutation confirms seizure locus SCN1A is also familial hemiplegic migraine locus FHM3. Pediatr. Neurol. 37, 407–410 (2007)

    Article  PubMed  Google Scholar 

  17. Cestèle, S., Schiavon, E., Rusconi, R., Franceschetti, S. & Mantegazza, M. Nonfunctional Nav1.1 familial hemiplegic migraine mutant transformed into gain of function by partial rescue of folding defects. Proc. Natl Acad. Sci. USA 110, 17546–17551 (2013)

    Article  PubMed  PubMed Central  ADS  CAS  Google Scholar 

  18. Priest, B. T. Future potential and status of selective sodium channel blockers for the treatment of pain. Curr. Opin. Drug Discov. Devel. 12, 682–692 (2009)

    CAS  PubMed  Google Scholar 

  19. Bohlen, C. J. & Julius, D. Receptor-targeting mechanisms of pain-causing toxins: How ow? Toxicon 60, 254–264 (2012)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  20. Olivera, B. M., Hillyard, D. R., Marsh, M. & Yoshikami, D. Combinatorial peptide libraries in drug design: lessons from venomous cone snails. Trends Biotechnol. 13, 422–426 (1995)

    Article  CAS  PubMed  Google Scholar 

  21. Black, J. A. et al. Spinal sensory neurons express multiple sodium channel α-subunit mRNAs. Brain Res. Mol. Brain Res. 43, 117–131 (1996)

    Article  CAS  PubMed  Google Scholar 

  22. McCormack, K. et al. Voltage sensor interaction site for selective small molecule inhibitors of voltage-gated sodium channels. Proc. Natl Acad. Sci. USA 110, E2724–E2732 (2013)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  23. Bosmans, F., Martin-Eauclaire, M. F. & Swartz, K. J. Deconstructing voltage sensor function and pharmacology in sodium channels. Nature 456, 202–208 (2008)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  24. Bosmans, F., Puopolo, M., Martin-Eauclaire, M. F., Bean, B. P. & Swartz, K. J. Functional properties and toxin pharmacology of a dorsal root ganglion sodium channel viewed through its voltage sensors. J. Gen. Physiol. 138, 59–72 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  25. Escoubas, P., Diochot, S., Célérier, M. L., Nakajima, T. & Lazdunski, M. Novel tarantula toxins for subtypes of voltage-dependent potassium channels in the Kv2 and Kv4 subfamilies. Mol. Pharmacol. 62, 48–57 (2002)

    Article  CAS  PubMed  Google Scholar 

  26. Smith, J. A. M., Davis, C. L. & Burgess, G. M. Prostaglandin E2-induced sensitization of bradykinin-evoked responses in rat dorsal root ganglion neurons is mediated by cAMP-dependent protein kinase A. Eur. J. Neurosci. 12, 3250–3258 (2000)

    Article  CAS  PubMed  Google Scholar 

  27. Wang, J. et al. Mapping the receptor site for α-scorpion toxins on a Na+ channel voltage sensor. Proc. Natl Acad. Sci. USA 108, 15426–15431 (2011)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  28. Bende, N. S. et al. A distinct sodium channel voltage-sensor locus determines insect selectivity of the spider toxin Dc1a. Nat. Commun. 5, 4350 (2014)

    Article  CAS  PubMed  ADS  Google Scholar 

  29. Zeitz, K. P. et al. The 5-HT3 subtype of serotonin receptor contributes to nociceptive processing via a novel subset of myelinated and unmyelinated nociceptors. J. Neurosci. 22, 1010–1019 (2002)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  30. Bautista, D. M. et al. The menthol receptor TRPM8 is the principal detector of environmental cold. Nature 448, 204–208 (2007)

    Article  CAS  PubMed  ADS  Google Scholar 

  31. Usoskin, D. et al. Unbiased classification of sensory neuron types by large-scale single-cell RNA sequencing. Nat. Neurosci. 18, 145–153 (2015)

    Article  CAS  PubMed  Google Scholar 

  32. Stucky, C. L., Medler, K. A. & Molliver, D. C. The P2Y agonist UTP activates cutaneous afferent fibers. Pain 109, 36–44 (2004)

    Article  CAS  PubMed  Google Scholar 

  33. Cavanaugh, D. J. et al. Distinct subsets of unmyelinated primary sensory fibers mediate behavioral responses to noxious thermal and mechanical stimuli. Proc. Natl Acad. Sci. USA 106, 9075–9080 (2009)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  34. Mishra, S. K., Tisel, S. M., Orestes, P., Bhangoo, S. K. & Hoon, M. A. TRPV1-lineage neurons are required for thermal sensation. EMBO J. 30, 582–593 (2011)

    Article  CAS  PubMed  Google Scholar 

  35. Mishra, S. K. & Hoon, M. A. Ablation of TrpV1 neurons reveals their selective role in thermal pain sensation. Mol. Cell. Neurosci. 43, 157–163 (2010)

    Article  CAS  PubMed  Google Scholar 

  36. Scherrer, G. et al. VGLUT2 expression in primary afferent neurons is essential for normal acute pain and injury-induced heat hypersensitivity. Proc. Natl Acad. Sci. USA 107, 22296–22301 (2010)

    Article  CAS  PubMed  PubMed Central  ADS  Google Scholar 

  37. Koltzenburg, M., Wall, P. D. & McMahon, S. B. Does the right side know what the left is doing? Trends Neurosci. 22, 122–127 (1999)

    Article  CAS  PubMed  Google Scholar 

  38. Kissin, I., Lee, S. S. & Bradley, E. L. Jr. Effect of prolonged nerve block on inflammatory hyperalgesia in rats: prevention of late hyperalgesia. Anesthesiology 88, 224–232 (1998)

    Article  CAS  PubMed  Google Scholar 

  39. Tsujino, H. et al. Activating transcription factor 3 (ATF3) induction by axotomy in sensory and motoneurons: A novel neuronal marker of nerve injury. Mol. Cell. Neurosci. 15, 170–182 (2000)

    Article  CAS  PubMed  Google Scholar 

  40. Brierley, S. M. & Linden, D. R. Neuroplasticity and dysfunction after gastrointestinal inflammation. Nat. Rev. Gastroenterol. Hepatol. 11, 611–627 (2014)

    Article  PubMed  Google Scholar 

  41. de Araujo, A. D. et al. Selenoether oxytocin analogues have analgesic properties in a mouse model of chronic abdominal pain. Nat. Commun. 5, 3165 (2014)

    Article  PubMed  ADS  CAS  Google Scholar 

  42. Jensen, H. S., Grunnet, M. & Bastlund, J. F. Therapeutic potential of Nav1.1 activators. Trends Pharmacol. Sci. 35, 113–118 (2014)

    Article  CAS  PubMed  Google Scholar 

  43. Khan, G. M., Chen, S. R. & Pan, H. L. Role of primary afferent nerves in allodynia caused by diabetic neuropathy in rats. Neuroscience 114, 291–299 (2002)

    Article  CAS  PubMed  Google Scholar 

  44. Tsunozaki, M. et al. A ‘toothache tree’ alkylamide inhibits Aδ mechanonociceptors to alleviate mechanical pain. J. Physiol. (Lond.) 591, 3325–3340 (2013)

    Article  CAS  Google Scholar 

  45. Pappagallo, M. Newer antiepileptic drugs: possible uses in the treatment of neuropathic pain and migraine. Clin. Ther. 25, 2506–2538 (2003)

    Article  CAS  PubMed  Google Scholar 

  46. Calabresi, P., Galletti, F., Rossi, C., Sarchielli, P. & Cupini, L. M. Antiepileptic drugs in migraine: from clinical aspects to cellular mechanisms. Trends Pharmacol. Sci. 28, 188–195 (2007)

    Article  CAS  PubMed  Google Scholar 

  47. Gilchrist, J. et al. Nav1.1 modulation by a novel triazole compound attenuates epileptic seizures in rodents. ACS Chem. Biol. 9, 1204–1212 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  48. Suter, M. R., Kirschmann, G., Laedermann, C. J., Abriel, H. & Decosterd, I. Rufinamide attenuates mechanical allodynia in a model of neuropathic pain in the mouse and stabilizes voltage-gated sodium channel inactivated state. Anesthesiology 118, 160–172 (2013)

    Article  CAS  PubMed  Google Scholar 

  49. Herzig, V. et al. ArachnoServer 2.0, an updated online resource for spider toxin sequences and structures. Nucleic Acids Res. 39, D653–D657 (2011)

    Article  CAS  PubMed  Google Scholar 

  50. Chow, C. Y., Cristofori-Armstrong, B., Undheim, E. A. B., King, G. F. & Rash, L. D. Three peptide modulators of the human voltage-gated sodium channel 1.7, an important analgesic target, from the venom of an Australian tarantula. Toxins (Basel) 7, 2494–2513 (2015)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  51. Kamber, B. et al. The synthesis of cystine peptides by iodine oxidation of S-trityl-cysteine and S-acetamidomethyl-cysteine peptides. Helv. Chim. Acta 63, 899–915 (1980)

    Article  CAS  Google Scholar 

  52. Vetter, I. et al. Isolation, characterization and total regioselective synthesis of the novel μO-conotoxin MfVIA from Conus magnificus that targets voltage-gated sodium channels. Biochem. Pharmacol. 84, 540–548 (2012)

    Article  CAS  PubMed  Google Scholar 

  53. Góngora-Benítez, M. et al. Acid-labile Cys-protecting groups for the Fmoc/tBu strategy: filling the gap. Org. Lett. 14, 5472–5475 (2012)

    Article  PubMed  CAS  Google Scholar 

  54. Frech, G. C., VanDongen, A. M., Schuster, G., Brown, A. M. & Joho, R. H. A novel potassium channel with delayed rectifier properties isolated from rat brain by expression cloning. Nature 340, 642–645 (1989)

    Article  CAS  PubMed  ADS  Google Scholar 

  55. Swartz, K. J. & MacKinnon, R. Hanatoxin modifies the gating of a voltage-dependent K+ channel through multiple binding sites. Neuron 18, 665–673 (1997)

    Article  CAS  PubMed  Google Scholar 

  56. Garcia, M. L., 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 33, 6834–6839 (1994)

    Article  CAS  PubMed  Google Scholar 

  57. Bohlen, C. J. et al. A bivalent tarantula toxin activates the capsaicin receptor, TRPV1, by targeting the outer pore domain. Cell 141, 834–845 (2010)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  58. Reeh, P. W. Sensory receptors in mammalian skin in an in vitro preparation. Neurosci. Lett. 66, 141–146 (1986)

    Article  CAS  PubMed  Google Scholar 

  59. Koltzenburg, M., Stucky, C. L. & Lewin, G. R. Receptive properties of mouse sensory neurons innervating hairy skin. J. Neurophysiol. 78, 1841–1850 (1997)

    Article  CAS  PubMed  Google Scholar 

  60. Brierley, S. M., Jones, R. C., III, Gebhart, G. F. & Blackshaw, L. A. Splanchnic and pelvic mechanosensory afferents signal different qualities of colonic stimuli in mice. Gastroenterology 127, 166–178 (2004)

    Article  PubMed  Google Scholar 

  61. Castro, J. et al. Linaclotide inhibits colonic nociceptors and relieves abdominal pain via guanylate cyclase-C and extracellular cyclic guanosine 3′,5′-monophosphate. Gastroenterology 145, 1334–1346 (2013)

    Article  CAS  PubMed  Google Scholar 

  62. de Araujo, A. D. et al. Selenoether oxytocin analogues have analgesic properties in a mouse model of chronic abdominal pain. Nat. Commun. 5, 3165 (2014)

    Article  PubMed  ADS  CAS  Google Scholar 

  63. Lewallen, K. A. et al. Assessing the role of the cadherin/catenin complex at the Schwann cell-axon interface and in the initiation of myelination. J. Neurosci. 31, 3032–3043 (2011)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  64. Bardoni, R. et al. Delta opioid receptors presynaptically regulate cutaneous mechanosensory neuron input to the spinal cord dorsal horn. Neuron 81, 1312–1327 (2014)

    Article  CAS  PubMed  PubMed Central  Google Scholar 

  65. Hughes, P. A. et al. Post-inflammatory colonic afferent sensitisation: different subtypes, different pathways and different time courses. Gut 58, 1333–1341 (2009)

    Article  CAS  PubMed  Google Scholar 

  66. Hughes, P. A. et al. Increased κ-opioid receptor expression and function during chronic visceral hypersensitivity. Gut 63, 1199–1200 (2014)

    Article  PubMed  Google Scholar 

Download references

Acknowledgements

We thank the Deutsche Arachnologische Gesellschaft and particularly I. Wendt, J. Broghammer, A. Schlosser, B. Rast, M. Luescher, C. and F. Schneider and H. Auer for providing arthropods for milking; W. Catterall for providing floxed Nav1.1 mice; and J. Poblete, J. Maddern, T. O’Donnell and A. Harrington for technical assistance. This work was supported by a T32 Postdoctoral Training Grant from the UCSF CVRI (J.D.O.), Ruth Kirschstein NIH postdoctoral (F32NS081907 to J.D.O.) and predoctoral (F31NS084646 to J.G. and F30DE023476 to J.J.E.) fellowships, the National Institutes of Health (R37NS065071 and R01NS081115 to D.J., R01NS091352 to F.B., R01NS040538 and R01NS070711 to C.L.S., and R37NS014627 and R01DA29204 to A.I.B.), the National Health and Medical Research Council of Australia (Project Grant APP1083480 to S.M.B., Program Grant APP1072113 and Principal Research Fellowship APP1044414 to G.F.K.), and a grant from the Wellcome Trust to A.I.B. S.M.B. is a NHMRC R.D Wright Biomedical Research Fellow.

Author information

Authors and Affiliations

Authors

Contributions

J.D.O., V.H., E.A.B.U., G.F.K. and D.J. carried out venom collection and screening, toxin purification and characterization. Z.D. and P.A. carried out Hm1a synthesis. J.D.O., J.G., C.Z., D.J. and F.B. designed, performed, and analysed electrophysiological and calcium imaging experiments to determine toxin mechanism and selectivity. J.J.E., J.D.O., X.W. A.I.B, and D.J. designed, performed and analysed histological experiments. A.D.W. and C.L.S. designed, performed, and analysed skin-nerve recordings. X.W., J.D.O., D.J., and A.I.B. designed, performed, and analysed behavioural experiments to assess somatic function. J.C., S.G.-C., L.G., G.Y.R. and S.M.B. designed, performed and analysed studies relating to colonic afferent and patch clamp pharmacological studies. All authors contributed to the discussion and interpretation of the results. J.D.O. and D.J. wrote the manuscript with contributions and suggestions from all authors.

Corresponding authors

Correspondence to Frank Bosmans, Glenn F. King or David Julius.

Ethics declarations

Competing interests

The authors’ institutions have submitted provisional patent applications based, in part, on the work described in this article.

Extended data figures and tables

Extended Data Figure 1 Hm1a and Hm1b selectively target Nav1.1 in sensory neurons.

a, Left, HPLC chromatogram showing reversed-phase C18 fractionation of Heteroscodra maculata venom; peaks containing Hm1a and Hm1b are labelled. Peptide sequences as determined by Edman degradation are displayed above. Middle, MALDI–TOF spectra of native undigested Hm1b (top) and native Hm1b digested with carboxypeptidase Y for 20 min (bottom), with inserted spectra showing the monoisotopic mass of each in daltons (Da). The observed mass difference of 146 Da between the intact and digested Hm1b corresponds to the final residue, Phe, with an amidated C terminus. Right, chromatograms show reversed-phase C18 HPLC profiles of native and synthetic Hm1a, which were indistinguishable when co-injected. b, Representative currents from oocytes expressing hNav subtypes before (black) and after (grey) bath application of ICA-121431 (500 nM). Currents were elicited during 1-Hz stimulation to induce use-dependent block. c, Top, amino acid sequence comparison of Hm1a with SGTx1, a related, but non-selective fast-inactivation inhibitor. Bottom, representative calcium imaging experiment comparing ICA-121431-mediated block of Hm1a- or SGTx1-evoked responses in cultured embryonic DRG neurons, with group data at right (**P < 0.01, ***P < 0.001, n = 4). d, Top, fraction of P0 mouse neurons responding to Hm1a versus SGTx1 (**P < 0.01). Bottom, ratiometric calcium responses elicited by SGTx1 (500 nM) in the presence and absence of ICA-121431 (500 nM). e, Dose–response curves for Hm1a inhibition of fast inactivation in oocytes expressing Nav1.1, Nav1.2 or Nav1.3. Sustained current at the end of a 100-ms pulse is normalized to peak current to quantify magnitude of the effect. EC50 values for hNav1.1 = 38 nM, hNav1.2 = 236 nM and hNav1.3 = 220 nM. f, Representative traces from oocytes expressing hNav subtypes in response to a saturating dose (on hNav1.1) of purified Hm1b during a 100-ms depolarization. g, rKv2.1 chimaeras containing different Nav1.9 S3b–S4 motifs were tested for sensitivity to hHm1a (100 nM). Representative traces (top) and summary data (bottom) show a lack of toxin sensitivity for each chimaera. h, Top left, representative currents from oocytes expressing mKv4.1 before (black) and after (red) bath application of Hm1a (5 μM). Middle, quantification of mKv4.1 blockade by synthetic or native Hm1a. Top right, comparison of sustained current during application of native or synthetic Hm1a (1 μM) shows similar effects on Nav1.1 inactivation. Bottom, representative traces (left) showing that outward currents in P0 trigeminal mouse neurons are unaffected by Hm1a (500 nM). Scatter plot (right, n = 10) shows no significant difference. i, Percentage of Hm1a (500 nM)-responsive neurons in various culture conditions as assessed by calcium imaging (n = 3–4,*P < 0.05). Error bars represent mean ± s.e.m. P values based on two-way ANOVA with post hoc Tukey’s test (c) or unpaired two-tailed Student’s t-test (d, i).

Extended Data Figure 2 Hm1a selectivity depends on DIV S1–S2 and S3b–S4 regions.

a, Top, alignments between Kv2.1 and hNav1.1 S3b–S4 regions from each domain (as indicated) with sequence of chimaeras shown below each alignment. Bottom, GV relationships from chimaeric channels expressed in oocytes in the absence (black) and presence (colours) of Hm1a (100 nM). b, Sequence alignment of hNav1.1 and rNav1.4 showing putative transmembrane segments (green) and regions swapped in chimaeric channels (grey). c, Top, using the background of Nav1.4 chimaera containing the S3b–S4 and S5–S6 regions of Nav1.1, individual residues were mutated in the S1–S2 loop to the cognate residue in Nav1.1. The D1376T and Y1379S point mutants in the chimaeric rNav1.4 channel reveal an increase in peak current after 100 nM Hm1a application (red) relative to untreated controls (black). Filled circles denote GV relationships, where oocytes were depolarized for 50 ms in 5-mV steps from a holding potential of −90 mV. Open circles denote steady-state inactivation (I/Imax) relationships, where oocytes were depolarized from −90 mV to +5 mV in 5-mV increments for 50 ms preceeding a 50-ms step to −15 mV. Middle, dot plot detailing per cent increase in peak conductance of each point mutant in response to 100 nM Hm1a treatment. Each point represents a single oocyte; red bars indicate 95% confidence interval. Mutations highlighted in orange (D1376T and Y1379S) are statistically different from S3b–S4/S5–S6 control (*P < 0.01, Student’s t-test). The Q1372E mutant did not generate currents. Bottom, alignment of DIV S1–S4 regions from relevant mouse and human Nav isoforms. Orange highlights location of residues in the S1–S2 loop that putatively contribute to the toxin effect. d, Left, stylized DIV with transmembrane segments represented as circles and extracellular loops as bars (black for native rNav1.4 channel and green for portions transplanted from hNav1.1). Middle, traces displaying effect of Hm1a on each chimaera depolarized to −15 mV from a holding potential of −90 mV. Right, conductance–voltage (G/Gmax) and steady-state inactivation (I/Imax) relationships of each channel and chimaera before and after toxin (black and red, respectively) across a voltage range spanning −90 mV to 0 mV from a holding potential of −90 mV in 5-mV increments. Scale bars as in Fig. 2. e, Dot plots displaying the effect of 100 nM Hm1a on peak current (left) and persistent current (right). Data in the left plot were generated by dividing peak conductance after Hm1a application by the peak conductance before. Right plot shows persistent current divided by peak current before (black) or after (red) toxin addition. Persistent current was determined by averaging current from the final millisecond of depolarization to 0 mV from a holding potential of −90 mV. Vertical bars indicate 95% confidence interval.

Extended Data Figure 3 Nav1.1 is expressed by myelinated, non-C-fibre sensory neurons.

a, Representative images showing expression of various molecular markers and their overlap with Nav1.1 transcripts. Markers include immunohistochemical staining for CGRP and tyrosine hydroxylase (TH) and in situ histochemistry for TRPM8 and 5-HT3 ion channel transcripts. Quantification of overlap for these markers is shown in Fig. 3. b, Quantification of the number of toxin-responsive cells in P0 mouse trigeminal cultures as assessed by calcium imaging (leftmost column) and the percentage of toxin-sensitive cells that responded to other agonists (1-(m-chlorophenyl)-biguanide (mCPBG), allyl isothiocyanate (AITC), capsaicin, and menthol activate 5-HT3, TRPA1, TRPV1 and TRPM8 channels, respectively), or bound the lectin IB4. c, Table including conduction velocity and Von Frey thresholds for skin-nerve experiments presented in Fig. 3d. Error bars represent mean ± s.e.m.

Extended Data Figure 4 Control experiments.

These data show control experiments related to Fig. 4. a, Representative DRG sections from peripherin-Cre adult mouse showing neurons that express Cre recombinase as visualized using a floxed-stop YFP reporter mouse. In situ hybridization histochemistry shows overlap with Nav1.1 transcripts (right). b, Quantification of overlap between YFP and Nav1.1. c, Comparison of ATF3 induction in DRG following sciatic nerve ligation (SNI) or intraplantar Hm1a injection. SNI induced ATF3 expression in >50% of DRG neurons whereas ATF3 induction after Hm1a injection was indistinguishable from vehicle (measured 1 or 3 days post-injection). d, Peripherin-Cre × floxed Nav1.1 mice were compared with wild-type littermates in the rotarod test. No significant differences were observed (unpaired Student’s t-test). Error bars represent mean ± s.e.m.

Extended Data Figure 5 A subset of colonic afferents does not express functional Nav1.1.

a, Left, representative ex vivo colonic single fibre recording from an Hm1a (100 nM)-non-responsive high-threshold fibre from a healthy mouse (arrows indicate application and removal of 2 g Von Frey hair stimulus). Middle, group data showing lack of Hm1a-mediated responses from a subset (9 out of 15) of fibres. Right, group data showing a population (5 out of 10) of healthy, high-threshold mechanoreceptor colonic afferents unaltered by ICA-121432 in the presence or absence of Hm1a (100 nM). b, Left, representative whole–cell current clamp recording of a retrogradely traced colonic DRG neuron in response to 500-ms current injection at rheobase. Recordings were made from the same neuron of a healthy control mouse before and after incubation with Hm1a (10 nM). Horizontal scale bar, 250 ms; vertical scale bar, 20 mV. Middle and right, group data show no effect of Hm1a application on electrical excitability in a sub-population (6 out of 11) of colonic DRG neurons. c, Left, representative high-threshold mechanoreceptive colonic fibres from CVH mice showing no change after application of Hm1a (100 nM). Middle, group data from Hm1a-non-responsive colonic fibres (4 out of 11). Right, group data showing a subpopulation of CVH colonic afferents (3 out of 10) unaltered by ICA-121432 in the presence or absence of Hm1a. d, Left, representative Hm1a-non-responsive colonic DRG neuron in whole-cell current clamp. Middle and right, group data show electrical excitability is unaltered by Hm1a in a subset (4 out of 11) of CVH colonic DRG neurons. Error bars represent mean ± s.e.m. No significant differences were observed using Student’s t-test (ad, middle; b, d, right) or one-way ANOVA with post hoc Bonferroni test (a, c, right).

Supplementary information

Supplementary Table

This file contains Supplementary Table 1, a list of venoms that produced no detectible or specific calcium response in cultured sensory neurons. (PDF 184 kb)

PowerPoint slides

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Osteen, J., Herzig, V., Gilchrist, J. et al. Selective spider toxins reveal a role for the Nav1.1 channel in mechanical pain. Nature 534, 494–499 (2016). https://doi.org/10.1038/nature17976

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1038/nature17976

This article is cited by

Comments

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.

Search

Quick links

Nature Briefing

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

Get the most important science stories of the day, free in your inbox. Sign up for Nature Briefing