Pharmacological characterisation of the highly NaV1.7 selective spider venom peptide Pn3a

Jennifer R Deuis, Zoltan Dekan, Joshua S Wingerd, Jennifer J Smith, Nehan R Munasinghe, Rebecca F Bhola, Wendy L Imlach, Volker Herzig, David A Armstrong, K Johan Rosengren, Frank Bosmans, Stephen G Waxman, Sulayman D Dib-Hajj, Pierre Escoubas, Michael S Minett, Macdonald J Christie, Glenn F King, Paul F Alewood, Richard J Lewis, John N Wood, Irina Vetter, Jennifer R Deuis, Zoltan Dekan, Joshua S Wingerd, Jennifer J Smith, Nehan R Munasinghe, Rebecca F Bhola, Wendy L Imlach, Volker Herzig, David A Armstrong, K Johan Rosengren, Frank Bosmans, Stephen G Waxman, Sulayman D Dib-Hajj, Pierre Escoubas, Michael S Minett, Macdonald J Christie, Glenn F King, Paul F Alewood, Richard J Lewis, John N Wood, Irina Vetter

Abstract

Human genetic studies have implicated the voltage-gated sodium channel NaV1.7 as a therapeutic target for the treatment of pain. A novel peptide, μ-theraphotoxin-Pn3a, isolated from venom of the tarantula Pamphobeteus nigricolor, potently inhibits NaV1.7 (IC50 0.9 nM) with at least 40-1000-fold selectivity over all other NaV subtypes. Despite on-target activity in small-diameter dorsal root ganglia, spinal slices, and in a mouse model of pain induced by NaV1.7 activation, Pn3a alone displayed no analgesic activity in formalin-, carrageenan- or FCA-induced pain in rodents when administered systemically. A broad lack of analgesic activity was also found for the selective NaV1.7 inhibitors PF-04856264 and phlotoxin 1. However, when administered with subtherapeutic doses of opioids or the enkephalinase inhibitor thiorphan, these subtype-selective NaV1.7 inhibitors produced profound analgesia. Our results suggest that in these inflammatory models, acute administration of peripherally restricted NaV1.7 inhibitors can only produce analgesia when administered in combination with an opioid.

Figures

Figure 1. Isolation of the novel spider…
Figure 1. Isolation of the novel spider peptide μ-TRTX-Pn3a from the venom of Pamphobeteus nigricolor.
(a) Photo of the male Pamphobeteus nigricolor specimen from which crude venom was obtained. (b) Chromatogram resulting from fractionation of the crude venom using RP-HPLC (purple dashed line indicates acetonitrile gradient). Corresponding activity of each fraction to inhibit veratridine-induced NaV1.3 responses is shown above (black circles). The arrow indicates the active peak. (c) Sequences of μ-TRTX-Pn3a and μ-TRTX-Pn3b identified by N-terminal sequencing.
Figure 2. Selectivity of μ-TRTX-Pn3a for hNa…
Figure 2. Selectivity of μ-TRTX-Pn3a for hNaV1.1-1.9 channels.
(a) Concentration-response curves and (b) comparative potency of Pn3a at hNaV1.1-1.9 assessed by whole-cell patch-clamp experiments. Pn3a most potently inhibited NaV1.7, with 40-fold selectivity over hNaV1.1, 100-fold selectivity over hNaV1.2, 1.3, 1.4 and 1.6, and 900-fold selectivity over NaV1.5, NaV1.8, and NaV1.9. Data are presented as mean ± SEM, with n = 3–9 cells per data point. (c) Representative hNaV1.1-1.9 current traces before (black) and after addition of Pn3a (red). Currents were obtained by a 20 ms pulse of −20 mV for NaV1.1-1.7, a 20 ms pulse +10 mV for NaV1.8, and a 40 ms pulse of −40 mV for NaV1.9. Pn3a (10 nM) selectively inhibited peak current at hNaV1.7 only.
Figure 3. Effect of μ-TRTX-Pn3a on the…
Figure 3. Effect of μ-TRTX-Pn3a on the electrophysiological parameters of hNaV1.1-1.8.
I-V curves before (black triangles) and after addition of Pn3a (white triangles). Pn3a (100 nM) inhibited peak current at all NaV subtypes but caused a rightward shift in the I-V curve at NaV1.7 only. G-V curves before (black circles) and after addition of Pn3a (white circles). Pn3a (100 nM) significantly shifted the V1/2 of voltage-dependence of activation to a more depolarized potential at NaV1.7 only (Δ + 21.3 mV). Voltage-dependence of steady-state fast inactivation curves before (black squares) and after addition of Pn3a (white squares). Pn3a (100 nM) caused small but significant shifts in the V1/2 of steady-state fast inactivation at NaV1.1, 1.2, 1.3, 1.4, 1.7 and 1.8. Data are presented as mean ± SEM, with n = 4–10 cells per data point.
Figure 4. Additional effects of μ-TRTX-Pn3a on…
Figure 4. Additional effects of μ-TRTX-Pn3a on the electrophysiological parameters of hNaV1.7.
(a) Time course of peak current reduction upon application of (arrow) 100 nM, 30 nM or 10 nM Pn3a at NaV1.7. Onset of block was measured with repetitive pulses to –20 mV every 20 s. Data are fitted with single-exponential functions and the time constant of block (τ) is indicated for each concentration. (b) Voltage-dependence of time-to-peak as a measure of activation kinetics at NaV1.7. Pn3a (100 nM) delayed time to peak between −20 mV and 15 mV. (c) Voltage-dependence of fast inactivation time constants at NaV1.7. Pn3a (100 nM) slowed the inactivation rate between −10 mV and 0 mV. (d) Rate of recovery from inactivation at NaV1.7. Pn3a (100 nM) slowed the rate of recovery from inactivation. Data are fitted with single-exponential functions and the time constant of recovery (τ) is indicated. (e) Steady-state slow inactivation at NaV1.7. Pn3a (100 nM) had no significant effect on the voltage-dependence of steady-state slow inactivation. Data are presented as mean ± SEM, with n = 4–6 cells per data point. Statistical significance was determined using two-way ANOVA, *P < 0.05 compared to control. (f) Representative ramp current elicited by a 50 ms depolarization from −100 to +20 mV at a rate of 2.4 mV/ms at NaV1.7. Pn3a (100 nM) inhibited ramp currents. (g) Potassium currents before (black) and in the presence of Pn3a (coloured) elicited by depolarisations to 70 mV in hNav1.7/rKv2.1 chimeras. Pn3a (300 nM) inhibited potassium currents in the DII hNaV1.7/KV2.1 and DIV hNaV1.7/KV2.1 chimeras. Scale bars: 50 ms (abscissa), 2 μA (DI, DII, DIV, KV2.1) and 1 μA (DIII) (ordinate axis).
Figure 5. Solution structure of Pn3a.
Figure 5. Solution structure of Pn3a.
(a) The family of the 20 lowest energy structures superimposed over the backbone with disulfide bonds shown in yellow (Protein Data Bank code 5T4R). (b) The lowest energy structure with β-strands shown as arrows and disulfide bonds in ball and stick representation.
Figure 6. Pn3a inhibits TTX-sensitive current in…
Figure 6. Pn3a inhibits TTX-sensitive current in small-diameter DRG neurons and eEPSC in C-fibres from lamina I neurons.
(a) Percentage inhibition of sodium current by TTX and Pn3a in IB4−, lightly IB4+ and strongly IB4+ small-diameter and large-diameter DRG neurons (n = 6–15). (b) Representative peak current vs time plot before and after addition of TTX and Pn3a in a small-diameter DRG neuron. (c) Representative peak current vs time plot before and after addition of TTX and Pn3a in a large-diameter DRG neuron (d) Representative current traces from a lamina I neuron indicating the position of the dorsal root evoked Aβ-, Aδ- and C-fibre currents pre- and post-treatment with Pn3a. (e) eEPSC amplitudes normalized to the baseline control for C-, Aδ-, Aβ-fibre currents in lamina I neurons treated with Pn3a (n = 7, 7 and 4, respectively). Data are presented as mean ± SEM. Statistical significance was determined using t-test compared to control, *P < 0.05.
Figure 7. Analgesic effects of selective Na…
Figure 7. Analgesic effects of selective NaV1.7 inhibitors.
(a) Pn3a (i.p.) dose-dependently reversed spontaneous pain behaviours elicited by intraplantar injection of the NaV1.7 activator OD1 in mice (over 10 min); n = 3–9 per group. (b) Time course of reversal of OD1-induced spontaneous pain behaviours by Pn3a (3 mg/kg i.p.); n = 6 per group. (c) Effect of Pn3a (3 mg/kg i.p.) and PF-04856264 (30 mg/kg i.p.) on noxious heat assessed using the hotplate test (50 °C) in mice; n = 6–16 per group. Dashed line represents time spent on the hotplate by NaV1.7Wnt knockout mice. (d) Time course of effect of Pn3a (3 mg/kg i.p.) and (e) PF-04856264 (30 mg/kg i.p.) on formalin-induced spontaneous pain behaviours in mice; n = 4–15 per group. Effect of Pn3a (3 mg/kg i.p.) and PF-04856264 (30 mg/kg i.p.) on (f) carrageenan-induced mechanical allodynia and (g) carrageenan-induced thermal allodynia in mice; n = 4–12 per group. Effect of Pn3a (0.3 nmoles i.t.) on (h) FCA-induced mechanical allodynia, (i) FCA-induced thermal allodynia, and (j) motor performance on the Rotarod in rats; n = 4–9 per group. Data are presented as mean ± SEM. Statistical significance was determined using t-test, one-way or two-way ANOVA with Dunnett’s post-test as appropriate, *P < 0.05 compared to control.
Figure 8. Selective Na V 1.7 inhibitors…
Figure 8. Selective NaV1.7 inhibitors synergize with opioids to produce analgesia.
(a) Time course of effect and (b) total pain behaviours with Pn3a (3 mg/kg i.p.), oxycodone (1 mg/kg i.p.) or the combination, in Phase I (0–10 min) and Phase II (10–55 min) of the formalin model in mice; n = 7–11 per group. Effect of Pn3a (3 mg/kg i.p.), PF-04856264 (30 mg/kg i.p.), oxycodone (0.66 mg/kg i.p.), or the combinations on (c) carrageenan-induced mechanical allodynia and (d) carrageenan-induced thermal allodynia in mice; n = 4–12 per group. (e) Effect of Phlotoxin-1 (50 μg/kg i.p.) and thiorphan (20 mg/kg i.p.) alone or in combination on acute heat responses; n = 6 per group (f) Effect of phlotoxin 1 (50 μg/kg i.p.) and buprenorphine (50 μg/kg i.p.) alone or in combination on acute heat responses; n = 6 per group. Effect of Pn3a (3 mg/kg i.p.) and gabapentin (100 mg/kg i.p.) alone or in combination on (g) carrageenan-induced mechanical allodynia and (h) carrageenan-induced thermal allodynia in mice; n = 3–8 per group. Data are presented as mean ± SEM. Statistical significance was determined using one-way or two-way ANOVA with Dunnett’s post-test as appropriate, *P < 0.05 compared to control.

References

    1. Catterall W. A., Goldin A. L. & Waxman S. G. International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol. Rev. 57, 397–409, doi: 10.1124/pr.57.4.4 (2005).
    1. Yu F. H. & Catterall W. A. Overview of the voltage-gated sodium channel family. Genome Biol. 4, 207 (2003).
    1. Goldberg Y. P. et al.. Loss-of-function mutations in the Nav1.7 gene underlie congenital indifference to pain in multiple human populations. Clinical genetics 71, 311–319, doi: 10.1111/j.1399-0004.2007.00790.x (2007).
    1. Cox J. J. et al.. An SCN9A channelopathy causes congenital inability to experience pain. Nature 444, 894–898, doi: 10.1038/nature05413 (2006).
    1. Fertleman C. R. et al.. SCN9A mutations in paroxysmal extreme pain disorder: allelic variants underlie distinct channel defects and phenotypes. Neuron 52, 767–774, doi: 10.1016/j.neuron.2006.10.006 (2006).
    1. Yang Y. et al.. Mutations in SCN9A, encoding a sodium channel alpha subunit, in patients with primary erythermalgia. J. Med. Genet. 41, 171–174 (2004).
    1. Cummins T. R., Dib-Hajj S. D. & Waxman S. G. Electrophysiological properties of mutant Nav1.7 sodium channels in a painful inherited neuropathy. J. Neurosci. 24, 8232–8236, doi: 10.1523/jneurosci.2695-04.2004 (2004).
    1. Trimmer J. S., Cooperman S. S., Agnew W. S. & Mandel G. Regulation of muscle sodium channel transcripts during development and in response to denervation. Dev. Biol. 142, 360–367 (1990).
    1. Rogart R. B., Cribbs L. L., Muglia L. K., Kephart D. D. & Kaiser M. W. Molecular cloning of a putative tetrodotoxin-resistant rat heart Na+ channel isoform. Proc. Natl. Acad. Sci. USA 86, 8170–8174 (1989).
    1. Caldwell J. H., Schaller K. L., Lasher R. S., Peles E. & Levinson S. R. Sodium channel Na(v)1.6 is localized at nodes of ranvier, dendrites, and synapses. Proc. Natl. Acad. Sci. USA 97, 5616–5620, doi: 10.1073/pnas.090034797 (2000).
    1. McCormack K. et al.. Voltage sensor interaction site for selective small molecule inhibitors of voltage-gated sodium channels. Proceedings of the National Academy of Sciences of the United States of America 110, E2724–2732, doi: 10.1073/pnas.1220844110 (2013).
    1. Ahuja S. et al.. Structural basis of Nav1.7 inhibition by an isoform-selective small-molecule antagonist. Science 350, aac5464, doi: 10.1126/science.aac5464 (2015).
    1. King G. F. & Hardy M. C. Spider-venom peptides: structure, pharmacology, and potential for control of insect pests. Annu. Rev. Entomol. 58, 475–496, doi: 10.1146/annurev-ento-120811-153650 (2013).
    1. Klint J. K. et al.. Spider-venom peptides that target voltage-gated sodium channels: pharmacological tools and potential therapeutic leads. Toxicon 60, 478–491, doi: 10.1016/j.toxicon.2012.04.337 (2012).
    1. Zlotkin E. The insect voltage-gated sodium channel as target of insecticides. Annu. Rev. Entomol. 44, 429–455, doi: 10.1146/annurev.ento.44.1.429 (1999).
    1. Loughney K., Kreber R. & Ganetzky B. Molecular analysis of the para locus, a sodium channel gene in Drosophila. Cell 58, 1143–1154 (1989).
    1. King G. F., Escoubas P. & Nicholson G. M. Peptide toxins that selectively target insect Na(V) and Ca(V) channels. Channels (Austin, Tex.) 2, 100–116 (2008).
    1. Bosmans F. & Swartz K. J. Targeting voltage sensors in sodium channels with spider toxins. Trends Pharmacol. Sci. 31, 175–182, doi: 10.1016/j.tips.2009.12.007 (2010).
    1. Deuis J. R. et al.. Analgesic Effects of GpTx-1, PF-04856264 and CNV1014802 in a Mouse Model of NaV1.7-Mediated Pain. Toxins (Basel) 8, doi: 10.3390/toxins8030078 (2016).
    1. Escoubas P. et al.. In 15th World Congress on Animal, Plant and Microbial Toxins 220–221 (Glasgow, Scotland, 2006).
    1. Minett M. S. et al.. Endogenous opioids contribute to insensitivity to pain in humans and mice lacking sodium channel Nav1.7. Nature communications 6, 8967, doi: 10.1038/ncomms9967 (2015).
    1. King G. F., Gentz M. C., Escoubas P. & Nicholson G. M. A rational nomenclature for naming peptide toxins from spiders and other venomous animals. Toxicon 52, 264–276, doi: 10.1016/j.toxicon.2008.05.020 (2008).
    1. Schmalhofer W. A. et al.. ProTx-II, a selective inhibitor of NaV1.7 sodium channels, blocks action potential propagation in nociceptors. Molecular pharmacology 74, 1476–1484, doi: 10.1124/mol.108.047670 (2008).
    1. Alexandrou A. J. et al.. Subtype-Selective Small Molecule Inhibitors Reveal a Fundamental Role for Nav1.7 in Nociceptor Electrogenesis, Axonal Conduction and Presynaptic Release. PLoS One 11, e0152405, doi: 10.1371/journal.pone.0152405 (2016).
    1. Dib-Hajj S. D., Yang Y., Black J. A. & Waxman S. G. The Na(V)1.7 sodium channel: from molecule to man. Nat. Rev. Neurosci. 14, 49–62, doi: 10.1038/nrn3404 (2013).
    1. Klint J. K. et al.. Seven novel modulators of the analgesic target NaV 1.7 uncovered using a high-throughput venom-based discovery approach. Br. J. Pharmacol. 172, 2445–2458, doi: 10.1111/bph.13081 (2015).
    1. Bosmans F., Martin-Eauclaire M. F. & Swartz K. J. Deconstructing voltage sensor function and pharmacology in sodium channels. Nature 456, 202–208, doi: 10.1038/nature07473 (2008).
    1. Brünger A. T. et al.. Crystallography & NMR system: A new software suite for macromolecular structure determination. Acta Crystallographica Section D: Biological Crystallography 54, 905–921 (1998).
    1. Takahashi H. et al.. Solution structure of hanatoxin1, a gating modifier of voltage-dependent K+ channels: common surface features of gating modifier toxins. Journal of molecular biology 297, 771–780 (2000).
    1. Jung H. J. et al.. Solution structure and lipid membrane partitioning of VSTx1, an inhibitor of the KvAP potassium channel. Biochemistry 44, 6015–6023 (2005).
    1. Minett M. S. et al.. Distinct Nav1.7-dependent pain sensations require different sets of sensory and sympathetic neurons. Nature communications 3, 791, doi: 10.1038/ncomms1795 (2012).
    1. Gingras J. et al.. Global Nav1.7 knockout mice recapitulate the phenotype of human congenital indifference to pain. PLoS One 9, e105895, doi: 10.1371/journal.pone.0105895 (2014).
    1. Nassar M. A. et al.. Nociceptor-specific gene deletion reveals a major role for Nav1.7 (PN1) in acute and inflammatory pain. Proceedings of the National Academy of Sciences of the United States of America 101, 12706–12711, doi: 10.1073/pnas.0404915101 (2004).
    1. Focken T. et al.. Discovery of Aryl Sulfonamides as Isoform-Selective Inhibitors of NaV1.7 with Efficacy in Rodent Pain Models. ACS Med. Chem. Lett. 7, 277–282, doi: 10.1021/acsmedchemlett.5b00447 (2016).
    1. Sun S., Cohen C. J. & Dehnhardt C. M. Inhibitors of voltage-gated sodium channel Nav1.7: patent applications since 2010. Pharmaceutical patent analyst 3, 509–521, doi: 10.4155/ppa.14.39 (2014).
    1. Deuis J. R. et al.. An animal model of oxaliplatin-induced cold allodynia reveals a crucial role for Nav1.6 in peripheral pain pathways. Pain 154, 1749–1757, doi: 10.1016/j.pain.2013.05.032 (2013).
    1. Deuis J. R. et al.. Analgesic effects of clinically used compounds in novel mouse models of polyneuropathy induced by oxaliplatin and cisplatin. Neuro-oncology 16, 1324–1332, doi: 10.1093/neuonc/nou048 (2014).
    1. Tanaka B. S. et al.. A gain-of-function mutation in Nav1.6 in a case of trigeminal neuralgia. Mol Med 22, doi: 10.2119/molmed.2016.00131 (2016).
    1. Black J. A., Frezel N., Dib-Hajj S. D. & Waxman S. G. Expression of Nav1.7 in DRG neurons extends from peripheral terminals in the skin to central preterminal branches and terminals in the dorsal horn. Molecular pain 8, 82, doi: 10.1186/1744-8069-8-82 (2012).
    1. Cummins T. R., Howe J. R. & Waxman S. G. Slow closed-state inactivation: a novel mechanism underlying ramp currents in cells expressing the hNE/PN1 sodium channel. J. Neurosci. 18, 9607–9619 (1998).
    1. Rush A. M., Cummins T. R. & Waxman S. G. Multiple sodium channels and their roles in electrogenesis within dorsal root ganglion neurons. The Journal of physiology 579, 1–14, doi: 10.1113/jphysiol.2006.121483 (2007).
    1. Deuis J. R. & Vetter I. The thermal probe test: A novel behavioral assay to quantify thermal paw withdrawal thresholds in mice. Temperature 3, 199–207, doi: 10.1080/23328940.2016.1157668 (2016).
    1. Fang X. et al.. Intense isolectin-B4 binding in rat dorsal root ganglion neurons distinguishes C-fiber nociceptors with broad action potentials and high Nav1.9 expression. J. Neurosci. 26, 7281–7292, doi: 10.1523/jneurosci.1072-06.2006 (2006).
    1. Fukuoka T. et al.. Comparative study of the distribution of the alpha-subunits of voltage-gated sodium channels in normal and axotomized rat dorsal root ganglion neurons. The Journal of comparative neurology 510, 188–206, doi: 10.1002/cne.21786 (2008).
    1. Fukuoka T. & Noguchi K. Comparative study of voltage-gated sodium channel alpha-subunits in non-overlapping four neuronal populations in the rat dorsal root ganglion. Neurosci. Res. 70, 164–171, doi: 10.1016/j.neures.2011.01.020 (2011).
    1. Black J. A. et al.. Spinal sensory neurons express multiple sodium channel alpha-subunit mRNAs. Brain Res. Mol. Brain Res. 43, 117–131 (1996).
    1. Weerasuriya A. & Mizisin A. P. The blood-nerve barrier: structure and functional significance. Methods Mol. Biol. 686, 149–173, doi: 10.1007/978-1-60761-938-3_6 (2011).
    1. Herzog R. I., Cummins T. R., Ghassemi F., Dib-Hajj S. D. & Waxman S. G. Distinct repriming and closed-state inactivation kinetics of Nav1.6 and Nav1.7 sodium channels in mouse spinal sensory neurons. The Journal of physiology 551, 741–750, doi: 10.1113/jphysiol.2003.047357 (2003).
    1. Renganathan M., Cummins T. R. & Waxman S. G. Contribution of Na(v)1.8 sodium channels to action potential electrogenesis in DRG neurons. J. Neurophysiol. 86, 629–640 (2001).
    1. Vasylyev D. V., Han C., Zhao P., Dib-Hajj S. & Waxman S. G. Dynamic-clamp analysis of wild-type human Nav1.7 and erythromelalgia mutant channel L858H. J. Neurophysiol. 111, 1429–1443, doi: 10.1152/jn.00763.2013 (2014).
    1. Emery E. C. et al.. Novel SCN9A mutations underlying extreme pain phenotypes: unexpected electrophysiological and clinical phenotype correlations. J. Neurosci. 35, 7674–7681, doi: 10.1523/jneurosci.3935-14.2015 (2015).
    1. Minett M. S. et al.. Pain without nociceptors? Nav1.7-independent pain mechanisms. Cell Rep 6, 301–312, doi: 10.1016/j.celrep.2013.12.033 (2014).
    1. Emery E. C., Luiz A. P. & Wood J. N. Nav1.7 and other voltage-gated sodium channels as drug targets for pain relief. Expert Opin. Ther. Targets 20, 975–983, doi: 10.1517/14728222.2016.1162295 (2016).
    1. Mambretti E. M. et al.. Functional and structural characterization of axonal opioid receptors as targets for analgesia. Molecular pain 12, doi: 10.1177/1744806916628734 (2016).
    1. Stein C. & Zollner C. Opioids and sensory nerves. Handb. Exp. Pharmacol. 495–518, doi: 10.1007/978-3-540-79090-7_14 (2009).
    1. Cao L. et al.. Pharmacological reversal of a pain phenotype in iPSC-derived sensory neurons and patients with inherited erythromelalgia. Sci Transl Med 8, 335ra356, doi: 10.1126/scitranslmed.aad7653 (2016).
    1. Herzig V. & Hodgson W. C. Neurotoxic and insecticidal properties of venom from the Australian theraphosid spider Selenotholus foelschei. Neurotoxicology 29, 471–475, doi: 10.1016/j.neuro.2008.03.002 (2008).
    1. Deuis J. R. et al.. Activation of kappa Opioid Receptors in Cutaneous Nerve Endings by Conorphin-1, a Novel Subtype-Selective Conopeptide, Does Not Mediate Peripheral Analgesia. ACS Chem. Neurosci. 6, 1751–1758, doi: 10.1021/acschemneuro.5b00113 (2015).
    1. Cummins T. R. et al.. Nav1.3 sodium channels: rapid repriming and slow closed-state inactivation display quantitative differences after expression in a mammalian cell line and in spinal sensory neurons. J. Neurosci. 21, 5952–5961 (2001).
    1. Vetter I. & Lewis R. J. Characterization of endogenous calcium responses in neuronal cell lines. Biochem. Pharmacol. 79, 908–920, doi: 10.1016/j.bcp.2009.10.020 (2010).
    1. Deuis J. R. et al.. Development of a muO-Conotoxin Analogue with Improved Lipid Membrane Interactions and Potency for the Analgesic Sodium Channel NaV1.8. J. Biol. Chem. 291, 11829–11842, doi: 10.1074/jbc.M116.721662 (2016).
    1. Güntert P. In Protein NMR Techniques 353–378 (Springer, 2004).
    1. Shen Y. & Bax A. Protein backbone and sidechain torsion angles predicted from NMR chemical shifts using artificial neural networks. Journal of biomolecular NMR 56, 227–241 (2013).
    1. Andersen N. H. et al.. Extracting information from the temperature gradients of polypeptide NH chemical shifts. 1. The importance of conformational averaging. Journal of the American Chemical Society 119, 8547–8561 (1997).
    1. Davis I. W. et al.. MolProbity: all-atom contacts and structure validation for proteins and nucleic acids. Nucleic acids research 35, W375–W383 (2007).
    1. Koradi R., Billeter M. & Wüthrich K. MOLMOL: a program for display and analysis of macromolecular structures. Journal of molecular graphics 14, 51–55 (1996).
    1. Murali S. S., Napier I. A., Rycroft B. K. & Christie M. J. Opioid-related (ORL1) receptors are enriched in a subpopulation of sensory neurons and prolonged activation produces no functional loss of surface N-type calcium channels. J. Physiol. 590, 1655–1667, doi: 10.1113/jphysiol.2012.228429 (2012).
    1. Imlach W. L., Bhola R. F., May L. T., Christopoulos A. & Christie M. J. A Positive Allosteric Modulator of the Adenosine A1 Receptor Selectively Inhibits Primary Afferent Synaptic Transmission in a Neuropathic Pain Model. Mol. Pharmacol. 88, 460–468, doi: 10.1124/mol.115.099499 (2015).
    1. Chaplan S. R., Bach F. W., Pogrel J. W., Chung J. M. & Yaksh T. L. Quantitative assessment of tactile allodynia in the rat paw. J. Neurosci. Methods 53, 55–63 (1994).

Source: PubMed

3
Subscribe