Hostname: page-component-8448b6f56d-c4f8m Total loading time: 0 Render date: 2024-04-18T07:16:42.236Z Has data issue: false hasContentIssue false

Sensory Neurons, Ion Channels, Inflammation and the Onset of Neuropathic Pain

Published online by Cambridge University Press:  02 December 2014

Patrick L. Stemkowski
Affiliation:
Centre for Neuroscience, University of Alberta, Edmonton Alberta, Canada. October 14, 2011 March 1, 2012
Peter A. Smith*
Affiliation:
Centre for Neuroscience, University of Alberta, Edmonton Alberta, Canada. October 14, 2011 March 1, 2012
*
Centre for Neuroscience and department of Pharmacology, 9-75 medical Sciences building, University of Alberta, Edmonton, Alberta, T6G 2h7, Canada. Email: peter.a.smith@ualberta.ca
Rights & Permissions [Opens in a new window]

Abstract

Core share and HTML view are not available for this content. However, as you have access to this content, a full PDF is available via the ‘Save PDF’ action button.

Neuropathic pain often fails to respond to conventional pain management procedures. here we review the aetiology of neuropathic pain as would result from peripheral neuropathy or injury. We show that inflammatory mediators released from damaged nerves and tissue are responsible for triggering ectopic activity in primary afferents and that this, in turn, provokes increased spinal cord activity and the development of ‘central sensitization’. Although evidence is mounting to support the role of interleukin-1β, prostaglandins and other cytokines in the onset of neuropathic pain, the clinical efficacy of drugs which antagonize or prevent the actions of these mediators is yet to be determined. basic science findings do, however, support the use of pre-emptive analgesia during procedures which involve nerve manipulation and the use of anti-inflammatory steroids as soon as possible following traumatic nerve injury.

Type
Review Article
Copyright
Copyright © The Canadian Journal of Neurological 2012

References

1.Merskey, H, Bogduk, N.Part III: Pain terms, a current list with definitions and notes on usage. In: Merskey, H, Bogduk, N, editors. Classification of chronic pain: descriptions of chronic pain syndromes and definitions of pain terms. Seattle: IASP Press; 1994. p. 20914.Google Scholar
2.Wang, H, Woolf, CJ.Pain TRPs. Neuron. 2005;46(1):912.CrossRefGoogle ScholarPubMed
3.Latremoliere, A, Woolf, CJ.Central sensitization: a generator of pain hypersensitivity by central neural plasticity. J Pain. 2009;10(9):895926.CrossRefGoogle Scholar
4.Cox, JJ, Reimann, F, Nicholas, AK, et al.An SCN9A channelopathy causes congenital inability to experience pain. Nature. 2006;444(7121):8948.CrossRefGoogle ScholarPubMed
5.Lee, MC, Mouraux, A, Iannetti, GD.Characterizing the cortical activity through which pain emerges from nociception. J Neurosci. 2009;29(24):790916.CrossRefGoogle ScholarPubMed
6.Wall, PD.Introduction to the fourth edition. In: Wall, PD, Melzack, R, editors. Textbook of pain. London: Harcourt Publishers Limited; 1999. p. 18.Google Scholar
7.Klit, H, Finnerup, NB, Jensen, TS.Central post-stroke pain: clinical characteristics, pathophysiology, and management. Lancet Neurol. 2009;8(9):85768.CrossRefGoogle ScholarPubMed
8.Price, DD.Psychological and neural mechanisms of the affective dimension of pain. Science. 2000;288(5472):176972.CrossRefGoogle ScholarPubMed
9.McCaffery, M, Pasero, C.Assessment: underlying complexities, misconceptions, and practical tools. In: McCaffery, M, Pasero, C, editors. Pain: clinical manual. 2 ed. St. Louis: Mosby; 1999. p. 35102.Google Scholar
10.Rainville, P, Carrier, B, Hofbauer, RK, et al.Dissociation of sensory and affective dimensions of pain using hypnotic modulation. Pain. 1999;82(2):15971.CrossRefGoogle ScholarPubMed
11.Rainville, P, Duncan, GH, Price, DD, et al.Pain affect encoded in human anterior cingulate but not somatosensory cortex. Science. 1997;277(5328):96871.CrossRefGoogle Scholar
12.Mogil, JS.Animal models of pain: progress and challenges. Nat Rev Neurosci. 2009;10(4):28394.CrossRefGoogle ScholarPubMed
13.Hummel, M, Lu, P, Cummons, TA, et al.The persistence of a long-term negative affective state following the induction of either acute or chronic pain. Pain. 2008;140(3):43645.CrossRefGoogle ScholarPubMed
14.Sandkuhler, J.Models and mechanisms of hyperalgesia and allodynia. Physiol Rev. 2009;89(2):70758.CrossRefGoogle ScholarPubMed
15.Backonja, MM, Stacey, B.Neuropathic pain symptoms relative to overall pain rating. J Pain. 2004;5(9):4917.CrossRefGoogle ScholarPubMed
16.Mogil, JS, Crager, SE.What should we be measuring in behavioral studies of chronic pain in animals? Pain. 2004;112(1-2):1215.CrossRefGoogle ScholarPubMed
17.Richardson, JD, Vasko, MR.Cellular mechanisms of neurogenic inflammation. J Pharmacol Exp Ther. 2002;302(3):83945.CrossRefGoogle ScholarPubMed
18.Irving, GA.Contemporary assessment and management of neuropathic pain. Neurology. 2005;64(12 Suppl 3):S217.CrossRefGoogle ScholarPubMed
19.Aley, KO, Messing, RO, Mochly-Rosen, D, et al.Chronic hypersensitivity for inflammatory nociceptor sensitization mediated by the epsilon isozyme of protein kinase C. J Neurosci. 2000;20(12):46805.CrossRefGoogle ScholarPubMed
20.Honore, P, Rogers, SD, Schwei, MJ, et al.Murine models of inflammatory, neuropathic and cancer pain each generates a unique set of neurochemical changes in the spinal cord and sensory neurons. Neuroscience. 2000;98(3):58598.CrossRefGoogle ScholarPubMed
21.Perl, ER.function of dorsal root ganglion neurons: an overview. In: Sheryl, A. Scott, editor. Sensory neurons diversity, development, and plasticity. New York: Oxford University Press; 1992. p. 323.Google Scholar
22.Gasser, HS.The control of excitation in the nervous system. Bull N Y Acad Med. 1937;13(6):32448.Google ScholarPubMed
23.Light, AR, Perl, ER.Unmyelinated afferent fibers are not only for pain anymore. J Comp Neurol. 2003;461(2):1379.CrossRefGoogle Scholar
24.Olausson, H, Lamarre, Y, Backlund, H, et al.Unmyelinated tactile afferents signal touch and project to insular cortex. Nat Neurosci. 2002;5(9):9004.CrossRefGoogle ScholarPubMed
25.Harper, AA, Lawson, SN.Conduction velocity is related to morphological cell type in rat dorsal root ganglion neurones. J Physiol. 1985;359:3146.CrossRefGoogle ScholarPubMed
26.Abdulla, FA, Smith, PA.Axotomy- and autotomy-induced changes in the excitability of rat dorsal root ganglion neurons. J Neurophysiol. 2001;85(2):63043.CrossRefGoogle Scholar
27.Aoki, Y, Takahashi, Y, Ohtori, S, et al.Distribution and immunocytochemical characterization of dorsal root ganglion neurons innervating the lumbar intervertebral disc in rats: a review. Life Sci. 2004;74(21):262742.CrossRefGoogle Scholar
28.Silverman, JD, Kruger, L.Selective neuronal glycoconjugate expression in sensory and autonomic ganglia: relation of lectin reactivity to peptide and enzyme markers. J Neurocytol. 1990;19(5):789801.CrossRefGoogle ScholarPubMed
29.Bradbury, EJ, Burnstock, G, McMahon, SB.The expression of P2X3 purinoreceptors in sensory neurons: effects of axotomy and glial-derived neurotrophic factor. Mol Cell Neurosci. 1998;12(4-5):25668.CrossRefGoogle ScholarPubMed
30.Lee, Y, Takami, K, Kawai, Y, et al.Distribution of calcitonin gene-related peptide in the rat peripheral nervous system with reference to its coexistence with substance P. Neuroscience. 1985;15(4):122737.CrossRefGoogle ScholarPubMed
31.Lee, Y, Kawai, Y, Shiosaka, S, et al.Coexistence of calcitonin gene-related peptide and substance P-like peptide in single cells of the trigeminal ganglion of the rat: immunohistochemical analysis. Brain Res. 1985;330(1):1946.CrossRefGoogle ScholarPubMed
32.Hokfelt, T, Elde, R, Johansson, O, et al.Immunohistochemical evidence for separate populations of somatostatin-containing and substance P-containing primary afferent neurons in the rat. Neuroscience. 1976;1(2):1316.CrossRefGoogle ScholarPubMed
33.Caterina, MJ, Schumacher, MA, Tominaga, M, et al.The capsaicin receptor: a heat-activated ion channel in the pain pathway. Nature. 1997;389(6653):81624.CrossRefGoogle ScholarPubMed
34.Guo, A, Vulchanova, L, Wang, J, et al.Immunocytochemical localization of the vanilloid receptor 1 (VR1): relationship to neuropeptides, the P2X3 purinoceptor and IB4 binding sites. Eur J Neurosci. 1999;11(3):94658.CrossRefGoogle ScholarPubMed
35.Aoki, YM, Ohtori, SM, Takahashi, KM, et al.Expression and co-expression of VR1, CGRP, and IB4-binding glycoprotein in dorsal root ganglion neurons in rats: differences between the disc afferents and the cutaneous afferents. Spine. 2005;30(13):1496500.CrossRefGoogle ScholarPubMed
36.Bennett, DLH, Michael, GJ, Ramachandran, N, et al.A distinct subgroup of small DRG cells express GDNF receptor components and GDNF is protective for these neurons after nerve injury. J Neurosci. 1998;18(8):305972.CrossRefGoogle ScholarPubMed
37.Bennett, DL, Averill, S, Clary, DO, et al.Postnatal changes in the expression of the trkA high-affinity NGF receptor in primary sensory neurons. Eur J Neurosci. 1996;8(10):22048.CrossRefGoogle ScholarPubMed
38.Stephens, HE, Belliveau, AC, Gupta, JS, et al.The role of neurotrophins in the maintenance of the spinal cord motor neurons and the dorsal root ganglia proprioceptive sensory neurons. Int J Dev Neurosci. 2005;23(7):61320.CrossRefGoogle ScholarPubMed
39.Lindsay, RM.The role of neurotrophic factors in functional maintenance of mature sensory neurons. In: Scott, SA, editor. Sensory neurons diversity, development, and plasticity. New York; 1992. p. 40420.Google Scholar
40.Averill, S, McMahon, SB, Clary, DO, et al.Immunocytochemical localization of trkA receptors in chemically identified subgroups of adult rat sensory neurons. Eur J Neurosci. 1995;7(7):148494.CrossRefGoogle ScholarPubMed
41.Molliver, DC, Wright, DE, Leitner, ML, et al.IB4-binding DRG neurons switch from NGF to GDNF dependence in early postnatal life. Neuron. 1997;19(4):84961.CrossRefGoogle ScholarPubMed
42.Airaksinen, MS, Saarma, M.The GDNF family: signalling, biological functions and therapeutic value. Nat Rev Neurosci. 2002;3(5):38394.CrossRefGoogle ScholarPubMed
43.Fjell, J, Cummins, TR, Dib-Hajj, SD, et al.Differential role of GDNF and NGF in the maintenance of two TTX-resistant sodium channels in adult DRG neurons. Brain Res Mol Brain Res. 1999;67(2):26782.CrossRefGoogle ScholarPubMed
44.Snider, WD, Wright, DE.Neurotrophins cause a new sensation. Neuron. 1996;16(2):22932.CrossRefGoogle Scholar
45.Luo, W, Wickramasinghe, SR, Savitt, JM, et al.A hierarchical NGF signaling cascade controls Ret-dependent and Ret-independent events during development of nonpeptidergic DRG neurons. Neuron. 2007;54(5):73954.CrossRefGoogle ScholarPubMed
46.Valdés-Sánchez, KM, Pérez-Villalba, A, Vega, JA, et al.BDNF is essentially required for the early postnatal survival of nociceptors. Dev Biol. 2010;339(2):46576.CrossRefGoogle ScholarPubMed
47.McMahon, SB, Armanini, MP, Ling, LH, et al.Expression and coexpression of Trk receptors in subpopulations of adult primary sensory neurons projecting to identified peripheral targets. Neuron. 1994;12(5):116171.CrossRefGoogle ScholarPubMed
48.Wright, DE, Snider, WD.Neurotrophin receptor mRNA expression defines distinct populations of neurons in rat dorsal root ganglia. J Comp Neurol. 1995;351(3):32938.CrossRefGoogle ScholarPubMed
49.Airaksinen, MS, Koltzenburg, M, Lewin, GR, et al.Specific subtypes of cutaneous mechanoreceptors require neurotrophin-3 following peripheral target innervation. Neuron. 1996;16(2):28795.CrossRefGoogle ScholarPubMed
50.Lumpkin, EA, Caterina, MJ.Mechanisms of sensory transduction in the skin. Nature. 2007;445(7130):85865.CrossRefGoogle ScholarPubMed
51.Voets, T, Droogmans, G, Wissenbach, U, et al.The principle of temperature-dependent gating in cold- and heat-sensitive TRP channels. Nature. 2004;430(7001):74854.CrossRefGoogle ScholarPubMed
52.Dhaka, A, Viswanath, V, Patapoutian, A.TRP ion channels and temperature sensation. Annu Rev Neurosci. 2006;29(1):13561.CrossRefGoogle ScholarPubMed
53.Ramsey, IS, Delling, M, Clapham, DE.An introduction to TRP channels. Annu Rev Physiol. 2006;68:61947.CrossRefGoogle ScholarPubMed
54.Schepers, RJ, Ringkamp, M.Thermoreceptors and thermosensitive afferents. Neurosci Biobehav Rev. 2009;33(3):20512.CrossRefGoogle ScholarPubMed
55.Caterina, MJ, Julius, D.The vanilloid receptor: a molecular gateway to the pain pathway. Annu Rev Neurosci. 2001;24:487517.CrossRefGoogle Scholar
56.Bandell, M, Story, GM, Hwang, SW, et al.Noxious cold ion channel TRPA1 is activated by pungent compounds and bradykinin. Neuron. 2004;41(6):84957.CrossRefGoogle Scholar
57.Peier, AM, Moqrich, A, Hergarden, AC, et al.A TRP channel that senses cold stimuli and menthol. Cell. 2002;108(5):70515.CrossRefGoogle Scholar
58.Kwan, KY, Allchorne, AJ, Vollrath, MA, et al.TRPA1 contributes to cold, mechanical, and chemical nociception but is not essential for hair-cell transduction. Neuron. 2006;50(2):27789.CrossRefGoogle Scholar
59.Suzuki, M, Mizuno, A, Kodaira, K, et al.Impaired pressure sensation in mice lacking TRPV4. J Biol Chem. 2003;278(25):226648.CrossRefGoogle ScholarPubMed
60.Alessandri-Haber, N, Joseph, E, Dina, OA, et al.TRPV4 mediates pain-related behavior induced by mild hypertonic stimuli in the presence of inflammatory mediator. Pain. 2005;118(1-2):709.CrossRefGoogle ScholarPubMed
61.Alloui, A, Zimmermann, K, Mamet, J, et al.TREK-1, a K+ channel involved in polymodal pain perception. EMBO J. 2006;25(11):236876.CrossRefGoogle Scholar
62.Maingret, F, Lauritzen, I, Patel, AJ, et al.TREK-1 is a heat-activated background K(+) channel. EMBO J. 2000;19(11):248391.CrossRefGoogle ScholarPubMed
63.Patel, AJ, Honore, E, Maingret, F, et al.A mammalian two pore domain mechano-gated S-like K+ channel. EMBO J. 1998;17(15):428390.CrossRefGoogle ScholarPubMed
64.Lumpkin, EA, Bautista, DM.Feeling the pressure in mammalian somatosensation. Curr Opin Neurobiol. 2005;15(4):3828.CrossRefGoogle ScholarPubMed
65.Morris, CE, Juranka, PF.Nav channel mechanosensitivity: activation and inactivation accelerate reversibly with stretch. Biophys J. 2007;93(3):82233.CrossRefGoogle ScholarPubMed
66.Vergnolle, N, Bunnett, NW, Sharkey, KA, et al.Proteinase-activated receptor-2 and hyperalgesia: a novel pain pathway. Nat Med. 2001;7(7):8216.CrossRefGoogle ScholarPubMed
67.Kruger, L, Perl, ER, Sedivec, MJ.Fine structure of myelinated mechanical nociceptor endings in cat hairy skin. J Comp Neurol. 1981;198(1):13754.CrossRefGoogle Scholar
68.Burgess, PR, Perl, ER.Myelinated afferent fibres responding specifically to noxious stimulation of the skin. J Physiol. 1967;190(3):54162.CrossRefGoogle ScholarPubMed
69.Bessou, P, Perl, ER.Response of cutaneous sensory units with unmyelinated fibers to noxious stimuli. J Neurophysiol. 1969;32(6):102543.CrossRefGoogle ScholarPubMed
70.Handwerker, HO, Kilo, S, Reeh, PW.Unresponsive afferent nerve fibres in the sural nerve of the rat. J Physiol. 1991;435:22942.CrossRefGoogle ScholarPubMed
71.Michaelis, M, Habler, HJ, Jaenig, W.Silent afferents: a separate class of primary afferents? Clin Exp Pharmacol Physiol. 1996;23(2):99105.CrossRefGoogle ScholarPubMed
72.Munger, BL, Ide, C.The structure and function of cutaneous sensory receptors. Arch Histol Cytol. 1988;51(1):134.CrossRefGoogle ScholarPubMed
73.Suzuki, M, Watanabe, Y, Oyama, Y, et al.Localization of mechanosensitive channel TRPV4 in mouse skin. Neurosci Lett. 2003;353(3):18992.CrossRefGoogle ScholarPubMed
74.Alvarez, FJ, Fyffe, RE.Nociceptors for the 21st century. Curr Rev Pain. 2000;4(6):4518.CrossRefGoogle Scholar
75.Denda, M, Nakatani, M, Ikeyama, K, et al.Epidermal keratinocytes as the forefront of the sensory system. Exp Dermatol. 2007;16(3):15761.CrossRefGoogle ScholarPubMed
76.González-Martínez, T, Germanà, P, Monjil, DF, et al.Absence of Meissner corpuscles in the digital pads of mice lacking functional TrkB. Brain Res. 2004;1002(1-2):1208.CrossRefGoogle ScholarPubMed
77.Johnson, KO.The roles and functions of cutaneous mechanoreceptors. Curr Opin Neurobiol. 2001;11(4):45561.CrossRefGoogle Scholar
78.Brodal, P.Peripheral parts of the somatosensory system. In: Brodal, P, editor. The central nervous system structure and function. New York: Oxford University Press; 2010. p. 16589.Google Scholar
79.Boulais, N, Misery, L.Merkel cells. J Am Acad Dermatol. 2007;57(1):14765.CrossRefGoogle Scholar
80.Matsuda, Y, Yoshida, S, Yonezawa, T.Tetrodotoxin sensitivity and Ca component of action potentials of mouse dorsal root ganglion cells cultured in vitro. Brain Res. 1978;154(1):6982.CrossRefGoogle ScholarPubMed
81.Yoshida, S, Matsuda, Y, Samejima, A.Tetrodotoxin-resistant sodium and calcium components of action potentials in dorsal root ganglion cells of the adult mouse. J Neurophysiol. 1978;41(5):1096106.CrossRefGoogle ScholarPubMed
82.Koerber, HR, Druzinsky, RE, Mendell, LM.Properties of somata of spinal dorsal root ganglion cells differ according to peripheral receptor innervated. J Neurophysiol. 1988;60(5):158496.CrossRefGoogle ScholarPubMed
83.Harper, AA, Lawson, SN.Electrical properties of rat dorsal root ganglion neurones with different peripheral nerve conduction velocities. J Physiol. 1985;359:4763.CrossRefGoogle ScholarPubMed
84.Gurtu, S, Smith, PA.Electrophysiological characteristics of hamster dorsal root ganglion cells and their response to axotomy. J Neurophysiol. 1988;59(2):40823.CrossRefGoogle ScholarPubMed
85.Nicholls, JG, Baylor, DA.Specific modalities and receptive fields of sensory neurons in CNS of the leech. J Neurophysiol. 1968;31(5):74056.CrossRefGoogle ScholarPubMed
86.Rose, RD, Koerber, HR, Sedivec, MJ, et al.Somal action potential duration differs in identified primary afferents. Neurosci Lett. 1986;63(3):25964.CrossRefGoogle ScholarPubMed
87.Koerber, HR, Mendell, LM.Functional heterogeneity of dorsal root ganglion cells. In: Scott, SA, editor. Sensory neurons diversity, development, and plasticity. New York: Oxford University Press; 1992. p. 7796.Google Scholar
88.Djouhri, L, Bleazard, L, Lawson, SN.Association of somatic action potential shape with sensory receptive properties in guinea-pig dorsal root ganglion neurones. J Physiol. 1998;513(Pt 3):85772.CrossRefGoogle ScholarPubMed
89.Fang, X, Djouhri, L, McMullan, S, 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. 2006;26(27):728192.CrossRefGoogle ScholarPubMed
90.Traub, RJ, Mendell, LM.The spinal projection of individual identified A-delta- and C-fibers. J Neurophysiol. 1988;59(1):4155.CrossRefGoogle ScholarPubMed
91.Nowycky, MC.Voltage-gated ion channels in dorsal root ganglion neurons. In: Scott, SA, editor. Sensory neurons diversity, development, and plasticity. New York: Oxford University Press; 1992. p. 97115.Google Scholar
92.Catterall, WA.From ionic currents to molecular mechanisms: the structure and function of voltage-gated sodium channels. Neuron. 2000;26(1):1325.CrossRefGoogle ScholarPubMed
93.Beneski, DA, Catterall, WA.Covalent labeling of protein components of the sodium channel with a photoactivable derivative of scorpion toxin. PNAS. 1980;77(1):63943.CrossRefGoogle ScholarPubMed
94.Catterall, WA, Goldin, AL, Waxman, SG.International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol Rev. 2005;57(4):397409.Google ScholarPubMed
95.Terlau, H, Heinemann, SH, Stühmer, W, et al.Mapping the site of block by tetrodotoxin and saxitoxin of sodium channel II. FEBS Lett. 1991;293(1-2):936.CrossRefGoogle ScholarPubMed
96.Caffrey, JM, Eng, DL, Black, JA, et al.Three types of sodium channels in adult rat dorsal root ganglion neurons. Brain Res. 1992;592(1-2):28397.CrossRefGoogle ScholarPubMed
97.Rizzo, MA, Kocsis, JD, Waxman, SG.Slow sodium conductances of dorsal root ganglion neurons: intraneuronal homogeneity and interneuronal heterogeneity. J Neurophysiol. 1994;72(6):2796815.CrossRefGoogle ScholarPubMed
98.Akopian, AN, Sivilotti, L, Wood, JN.A tetrodotoxin-resistant voltage-gated sodium channel expressed by sensory neurons. Nature. 1996;379(6562):25762.CrossRefGoogle ScholarPubMed
99.Sangameswaran, L, Delgado, SG, Fish, LM, et al.Structure and function of a novel voltage-gated, tetrodotoxin-resistant sodium channel specific to sensory neurons. J Biol Chem. 1996;271(11):59536.CrossRefGoogle ScholarPubMed
100.Dib-Hajj, SD, Tyrrell, L, Black, JA, et al.NaN, a novel voltage-gated Na channel, is expressed preferentially in peripheral sensory neurons and down-regulated after axotomy. PNAS. 1998;95(15):89638.CrossRefGoogle Scholar
101.Rush, AM, Cummins, TR, Waxman, SG.multiple sodium channels and their roles in electrogenesis within dorsal root ganglion neurons. J Physiol. 2007;579(Pt 1):114.CrossRefGoogle ScholarPubMed
102.Abdulla, FA, Smith, PA.Changes in Na(+) channel currents of rat dorsal root ganglion neurons following axotomy and axotomy-induced autotomy. J Neurophysiol. 2002;88(5):251829.CrossRefGoogle Scholar
103.Cummins, TR, Dib-Hajj, SD, Black, JA, et al.A novel persistent tetrodotoxin-resistant sodium current in SNS-null and wild-type small primary sensory neurons. J Neurosci. 1999;19(24):43RC.CrossRefGoogle ScholarPubMed
104.Gover, TD, Moreira, TH, Weinreich, D.Role of calcium in regulating primary sensory neuronal excitability. In: Canning, BJ, Spina, D, editors. Sensory nerves: handbook of experimental pharmacology. Berlin Heidelberg: Springer; 2009. p. 56387.CrossRefGoogle Scholar
105.Catterall, WA, Perez-Reyes, E, Snutch, TP, et al.International Union of Pharmacology. XLVIII. Nomenclature and structure-function relationships of voltage-gated calcium channels. Pharmacol Rev. 2005;57(4):41125.Google ScholarPubMed
106.Catterall, WA.Structure and regulation of voltage-gated Ca2+ channels. Annu Rev Cell Dev Biol. 2000;16(1):52155.CrossRefGoogle ScholarPubMed
107.Scroggs, RS, Fox, AP.Calcium current variation between acutely isolated adult rat dorsal root ganglion neurons of different size. J Physiol. 1992;445:63958.CrossRefGoogle ScholarPubMed
108.Fuchs, A, Rigaud, M, Sarantopoulos, CD, et al.Contribution of calcium channel subtypes to the intracellular calcium signal in sensory neurons: the effect of injury. Anesthesiology. 2007;107(1):11727.CrossRefGoogle Scholar
109.Acosta, CG, Lopez, HS.δ Opioid receptor modulation of several voltage-dependent Ca2+ currents in rat sensory neurons. J Neurosci. 1999;19(19):833748.CrossRefGoogle ScholarPubMed
110.Schroeder, JE, Fischbach, PS, McCleskey, EW.T-type calcium channels: heterogeneous expression in rat sensory neurons and selective modulation by phorbol esters. J Neurosci. 1990;10(3):94751.CrossRefGoogle ScholarPubMed
111.Lu, SG, Zhang, X, Gold, MS.Intracellular calcium regulation among subpopulations of rat dorsal root ganglion neurons. J Physiol. 2006;577(1):16990.CrossRefGoogle ScholarPubMed
112.Abdulla, FA, Smith, PA.Axotomy- and autotomy-induced changes in Ca2+and K+ channel currents of rat dorsal root ganglion neurons. J Neurophysiol. 2001;85(2):64458.CrossRefGoogle ScholarPubMed
113.Zamponi, GW, Lewis, RJ, Todorovic, SM, et al.Role of voltage-gated calcium channels in ascending pain pathways. Brain Res Rev. 2009;60(1):849.CrossRefGoogle ScholarPubMed
114.Todorovic, SM, Jevtovic-Todorovic, V.The role of T-type calcium channels in peripheral and central pain processing. CNS Neurol Disord Drug Targets. 2006;5(6):63953.CrossRefGoogle ScholarPubMed
115.Felix, R, Gurnett, CA, De Waard, M, et al.Dissection of functional domains of the voltage-dependent Ca2+ channel α2δ subunit. J Neurosci. 1997;17(18):688491.CrossRefGoogle ScholarPubMed
116.Newton, RA, Bingham, S, Case, PC, et al.Dorsal root ganglion neurons show increased expression of the calcium channel [alpha]2[delta]-1 subunit following partial sciatic nerve injury. Mol Brain Res. 2001;95(1-2):18.CrossRefGoogle Scholar
117.Luo, ZD, Chaplan, SR, Higuera, ES, et al.Upregulation of dorsal root ganglion [alpha]2[delta] calcium channel subunit and its correlation with allodynia in spinal nerve-injured rats. J Neurosci. 2001;21(6):186875.CrossRefGoogle ScholarPubMed
118.Boroujerdi, A, Kim, HK, Lyu, YS, et al.Injury discharges regulate calcium channel alpha-2-delta-1 subunit upregulation in the dorsal horn that contributes to initiation of neuropathic pain. Pain. 2008;139(2):35866.CrossRefGoogle ScholarPubMed
119.Luo, ZD, Calcutt, NA, Higuera, ES, et al.Injury type-specific calcium channel α2δ-1 subunit up-regulation in rat neuropathic pain models correlates with antiallodynic effects of gabapentin. J Pharmacol Exp Ther. 2002;303(3):1199205.CrossRefGoogle ScholarPubMed
120.Field, MJ, Cox, PJ, Stott, E, et al.Identification of the α2-δ-1 subunit of voltage-dependent calcium channels as a molecular target for pain mediating the analgesic actions of pregabalin. PNAS. 2006;103(46):1753742.CrossRefGoogle ScholarPubMed
121.Marais, E, Klugbauer, N, Hofmann, F.Calcium channel α2δ subunits-structure and gabapentin binding. Mol Pharmacol. 2001;59(5):12438.CrossRefGoogle ScholarPubMed
122.Stefani, A, Spadoni, F, Bernardi, G.Gabapentin inhibits calcium currents in isolated rat brain neurons. Neuropharmacology. 1998;37(1):8391.CrossRefGoogle ScholarPubMed
123.Gee, NS, Brown, JP, Dissanayake, VUK, et al.The novel anticonvulsant drug, gabapentin (Neurontin), binds to the subunit of a calcium channel. J Biol Chem. 1996;271(10):576876.CrossRefGoogle Scholar
124.Gutman, GA, Chandy, KG, Grissmer, S, et al.International Union of Pharmacology. LIII. Nomenclature and molecular relationships of voltage-gated potassium channels. Pharmacol Rev. 2005;57(4):473508.CrossRefGoogle ScholarPubMed
125.Ocaña, M, Cendán, CM, Cobos, EJ, et al.Potassium channels and pain: present realities and future opportunities. Eur J Pharmacol. 2004;500(1-3):20319.CrossRefGoogle ScholarPubMed
126.Bayliss, DA, Barrett, PQ.Emerging roles for two-pore-domain potassium channels and their potential therapeutic impact. Trends Pharmacol Sci. 2008;29(11):56675.CrossRefGoogle ScholarPubMed
127.Lesage, F, Lazdunski, M.Molecular and functional properties of two-pore-domain potassium channels. Am J Physiol Renal Physiol. 2000;279(5):F793801.CrossRefGoogle ScholarPubMed
128.Hibino, H, Inanobe, A, Furutani, K, et al.Inwardly rectifying potassium channels: their structure, function, and physiological roles. Physiol Rev. 2010;90(1):291366.CrossRefGoogle ScholarPubMed
129.Pischalnikova, A, Sokolova, O.The domain and conformational organization in potassium voltage-gated ion channels. J Neuroimmune Pharmacol. 2009;4(1):7182.CrossRefGoogle ScholarPubMed
130.Ghatta, S, Nimmagadda, D, Xu, X, et al.Large-conductance, calcium-activated potassium channels: structural and functional implications. Pharmacol Ther. 2006;110(1):10316.CrossRefGoogle ScholarPubMed
131.Kostyuk, PG, Veselovsky, NS, Fedulova, SA, et al.Ionic currents in the somatic membrane of rat dorsal root ganglion neurons-III. Potassium currents. Neuroscience. 1981;6(12):243944.Google ScholarPubMed
132.Kameyama, M.Ionic currents in cultured dorsal root ganglion cells from adult guinea pigs. J Membr Biol. 1983;72(3):195203.CrossRefGoogle ScholarPubMed
133.Kasai, H, Kameyama, M, Yamaguchi, K, et al.Single transient K channels in mammalian sensory neurons. Biophys J. 1986;49(6):12437.CrossRefGoogle ScholarPubMed
134.Penner, R, Petersen, M, Pierau, FK, et al.Dendrotoxin: a selective blocker of a non-inactivating potassium current in guinea-pig dorsal root ganglion neurones. Pflugers Arch. 1986;407(4):3659.CrossRefGoogle ScholarPubMed
135.Stansfeld, CE, Marsh, SJ, Halliwell, JV, et al.4-Aminopyridine and dendrotoxin induce repetitive firing in rat visceral sensory neurones by blocking a slowly inactivating outward current. Neurosci Lett. 1986;64(3):299304.CrossRefGoogle ScholarPubMed
136.Stansfeld, C, Feltz, A.Dendrotoxin-sensitive K+ channels in dorsal root ganglion cells. Neurosci Lett. 1988;93(1):4955.CrossRefGoogle ScholarPubMed
137.McFarlane, S, Cooper, E.Kinetics and voltage dependence of A-type currents on neonatal rat sensory neurons. J Neurophysiol. 1991;66(4):138091.CrossRefGoogle ScholarPubMed
138.Gold, MS, Shuster, MJ, Levine, JD.Characterization of six voltage-gated K+ currents in adult rat sensory neurons. J Neurophysiol. 1996;75(6):262946.CrossRefGoogle ScholarPubMed
139.Yoshida, S, Matsumoto, S.Effects of α-Dendrotoxin on K+ currents and action potentials in tetrodotoxin-resistant adult rat trigeminal ganglion neurons. J Pharmacol Exp Ther. 2005;314(1):43745.CrossRefGoogle Scholar
140.Everill, B, Rizzo, MA, Kocsis, JD.Morphologically identified cutaneous afferent DRG neurons express three different potassium currents in varying proportions. J Neurophysiol. 1998;79(4):181424.CrossRefGoogle ScholarPubMed
141.Vydyanathan, A, Wu, ZZ, Chen, SR, et al.A-type voltage-gated K+ currents influence firing properties of isolectin B4-positive but not isolectin B4-negative primary sensory neurons. J Neurophysiol. 2005;93(6):34019.CrossRefGoogle Scholar
142.Villiere, V, McLachlan, EM.Electrophysiological properties of neurons in intact rat dorsal root ganglia classified by conduction velocity and action potential duration. J Neurophysiol. 1996;76(3):192441.CrossRefGoogle ScholarPubMed
143.Connor, JA, Stevens, CF.Voltage clamp studies of a transient outward membrane current in gastropod neural somata. J Physiol. 1971;213(1):2130.CrossRefGoogle ScholarPubMed
144.Catacuzzeno, L, Fioretti, B, Pietrobon, D, et al.The differential expression of low-threshold K+ currents generates distinct firing patterns in different subtypes of adult mouse trigeminal ganglion neurones. J Physiol. 2008;586(Pt 21):510118.CrossRefGoogle ScholarPubMed
145.Gold, MS, Shuster, MJ, Levine, JD.Role of a Ca(2+)-dependent slow afterhyperpolarization in prostaglandin E2-induced sensitization of cultured rat sensory neurons. Neurosci Lett. 1996;205(3):1614.CrossRefGoogle Scholar
146.Scholz, A, Gruss, M, Vogel, W.Properties and functions of calcium-activated K+ channels in small neurones of rat dorsal root ganglion studied in a thin slice preparation. J Physiol. 1998;513(Pt 1):5569.CrossRefGoogle Scholar
147.Zhang, XL, Mok, LP, Katz, EJ, et al.BKCa currents are enriched in a subpopulation of adult rat cutaneous nociceptive dorsal root ganglion neurons. Eur J Neurosci. 2010;31(3):45062.CrossRefGoogle Scholar
148.Christian, EP, Togo, J, Naper, KE.Guinea pig visceral C-fiber neurons are diverse with respect to the K+ currents involved in action-potential repolarization. J Neurophysiol. 1994;71(2):56174.CrossRefGoogle Scholar
149.Morita, K, Katayama, Y.Calcium-dependent slow outward current in visceral primary afferent neurones of the rabbit. Pflugers Arch. 1989;414(2):1717.CrossRefGoogle ScholarPubMed
150.Bahia, PK, Suzuki, R, Benton, DC, et al.A functional role for small-conductance calcium-activated potassium channels in sensory pathways including nociceptive processes. J Neurosci. 2005;25(14):348998.CrossRefGoogle ScholarPubMed
151.Kaupp, UB, Seifert, R.Molecular diversity of pacemaker ion channels. Annu Rev Physiol. 2001;63:23557.CrossRefGoogle ScholarPubMed
152.Santoro, B, Liu, DT, Yao, H, et al.Identification of a gene encoding a hyperpolarization-activated pacemaker channel of brain. Cell. 1998;93(5):71729.CrossRefGoogle ScholarPubMed
153.Ulens, C, Tytgat, J.Functional heteromerization of HCN1 and HCN2 pacemaker channels. J Biol Chem. 2001;276(9):606972.CrossRefGoogle ScholarPubMed
154.Orio, P, Madrid, R, de la Peña, E, et al.Characteristics and physiological role of hyperpolarization activated currents in mouse cold thermoreceptors. J Physiol. 2009;587(9):196176.CrossRefGoogle ScholarPubMed
155.Mayer, ML, Westbrook, GL.A voltage-clamp analysis of inward (anomalous) rectification in mouse spinal sensory ganglion neurones. J Physiol. 1983;340:1945.CrossRefGoogle ScholarPubMed
156.Scroggs, RS, Todorovic, SM, Anderson, EG, et al.Variation in IH, IIR, and ILEAK between acutely isolated adult rat dorsal root ganglion neurons of different size. J Neurophysiol. 1994;71(1):2719.CrossRefGoogle ScholarPubMed
157.Pape, HC.Queer current and pacemaker: the hyperpolarizationactivated cation current in neurons. Annu Rev Physiol. 1996;58:299327.CrossRefGoogle ScholarPubMed
158.Cho, Hj, Staikopoulos, V, Ivanusic, J, et al.Hyperpolarization-activated cyclic-nucleotide gated 4 (HCN4) protein is expressed in a subset of rat dorsal root and trigeminal ganglion neurons. Cell Tissue Res. 2009;338(2):1717.CrossRefGoogle Scholar
159.Jiang, YQ, Xing, GG, Wang, SL, et al.Axonal accumulation of hyperpolarization-activated cyclic nucleotide-gated cation channels contributes to mechanical allodynia after peripheral nerve injury in rat. Pain. 2008;137(3):495506.CrossRefGoogle ScholarPubMed
160.Tu, H, Deng, L, Sun, Q, et al.Hyperpolarization-activated, cyclic nucleotide-gated cation channels: roles in the differential electrophysiological properties of rat primary afferent neurons. J Neurosci Res. 2004;76(5):71322.CrossRefGoogle ScholarPubMed
161.Kouranova, EV, Strassle, BW, Ring, RH, et al.Hyperpolarization-activated cyclic nucleotide-gated channel mRNA and protein expression in large versus small diameter dorsal root ganglion neurons: correlation with hyperpolarization-activated current gating. Neuroscience. 2008;153(4):100819.CrossRefGoogle ScholarPubMed
162.Momin, A, Cadiou, H, Mason, A, et al.Role of the hyperpolarization-activated current Ih in somatosensory neurons. J Physiol. 2008;586(Pt 24):591129.CrossRefGoogle ScholarPubMed
163.Viana, F, Belmonte, C.Funny currents are becoming serious players in nociceptor’s sensitization. J Physiol. 2008;586(Pt 24):58412.CrossRefGoogle ScholarPubMed
164.Doubell, TP, Mannion, JR, Woolf, CJ.The dorsal horn: state-dependent sensory processing, plasticity and the generation of pain. In: Wall, PD, Melzack, R, editors. Textbook of pain. London: Harcourt Publishers Limited; 1999. p. 16581.Google Scholar
165.Rexed, B.A cytoarchitectonic atlas of the spinal cord in the cat. J Comp Neurol. 1954;100(2):297379.CrossRefGoogle ScholarPubMed
166.Rexed, B.The cytoarchitectonic organization of the spinal cord in the cat. J Comp Neurol. 1952;96(3):41495.CrossRefGoogle ScholarPubMed
167.Brown, AG, Rose, PK, Snow, PJ.The morphology of hair follicle afferent fibre collaterals in the spinal cord of the cat. J Physiol. 1977;272(3):77997.CrossRefGoogle ScholarPubMed
168.Brown, AG, Rose, PK, Snow, PJ.Morphology and organization of axon collaterals from afferent fibres of slowly adapting type I units in cat spinal cord. J Physiol. 1978;277:1527.CrossRefGoogle ScholarPubMed
169.Brown, AG, Fyffe, RE, Noble, R.Projections from Pacinian corpuscles and rapidly adapting mechanoreceptors of glabrous skin to the cat’s spinal cord. J Physiol. 1980;307:385400.CrossRefGoogle Scholar
170.Brown, AG, Fyffe, RE, Rose, PK, et al.Spinal cord collaterals from axons of type II slowly adapting units in the cat. J Physiol. 1981;316:46980.CrossRefGoogle ScholarPubMed
171.Woolf, CJ.Central terminations of cutaneous mechanoreceptive afferents in the rat lumbar spinal cord. J Comp Neurol. 1987;261(1):10519.CrossRefGoogle ScholarPubMed
172.Light, AR, Trevino, DL, Perl, ER.Morphological features of functionally defined neurons in the marginal zone and substantia gelatinosa of the spinal dorsal horn. J Comp Neurol. 1979;186(2):15171.CrossRefGoogle ScholarPubMed
173.Sugiura, Y, Terui, N, Hosoya, Y.Difference in distribution of central terminals between visceral and somatic unmyelinated (C) primary afferent fibers. J Neurophysiol. 1989;62(4):83440.CrossRefGoogle ScholarPubMed
174.Traub, RJ, Solodkin, A, Ruda, MA.Calcitonin gene-related peptide immunoreactivity in the cat lumbosacral spinal cord and the effects of multiple dorsal rhizotomies. J Comp Neurol. 1989;287(2):22537.CrossRefGoogle ScholarPubMed
175.Ribeiro-da-Silva, A, Cuello, AC.Organization of peptidergic neurons in the dorsal horn of the spinal cord: anatomical and functional correlates. Prog Brain Res. 1995;104:4159.CrossRefGoogle ScholarPubMed
176.Light, AR, Perl, ER.Spinal termination of functionally identified primary afferent neurons with slowly conducting myelinated fibers. J Comp Neurol. 1979;186(2):13350.CrossRefGoogle Scholar
177.Woolf, CJ.Evidence for a central component of post-injury pain hypersensitivity. Nature. 1983;306(5944):6868.CrossRefGoogle ScholarPubMed
178.Govrin-Lippmann, R, Devor, M.Ongoing activity in severed nerves: source and variation with time. Brain Res. 1978;159(2):40610.CrossRefGoogle ScholarPubMed
179.Wall, PD, Devor, M.Sensory afferent impulses originate from dorsal root ganglia as well as from the periphery in normal and nerve injured rats. Pain. 1983;17(4):32139.CrossRefGoogle ScholarPubMed
180.Kajander, KC, Wakisaka, S, Bennett, GJ.Spontaneous discharge originates in the dorsal root ganglion at the onset of a painful peripheral neuropathy in the rat. Neurosci Lett. 1992;138(2):2258.CrossRefGoogle ScholarPubMed
181.Kim, YI, Na, HS, Kim, SH, et al.Cell type-specific changes of the membrane properties of peripherally-axotomized dorsal root ganglion neurons in a rat model of neuropathic pain. Neuroscience. 1998;86(1):3019.CrossRefGoogle Scholar
182.Zhang, JM, Donnelly, DF, Song, XJ, et al.Axotomy increases the excitability of dorsal root ganglion cells with unmyelinated axons. J Neurophysiol. 1997;78(5):27904.CrossRefGoogle ScholarPubMed
183.Study, RE, Kral, MG.Spontaneous action potential activity in isolated dorsal root ganglion neurons from rats with a painful neuropathy. Pain. 1996;65(2-3):23542.CrossRefGoogle ScholarPubMed
184.Ma, C, LaMotte, RH.Enhanced excitability of dissociated primary sensory neurons after chronic compression of the dorsal root ganglion in the rat. Pain. 2005;113(1-2):10612.CrossRefGoogle ScholarPubMed
185.Everill, B, Cummins, TR, Waxman, SG, et al.Sodium currents of large (Aβ-type) adult cutaneous afferent dorsal root ganglion neurons display rapid recovery from inactivation before and after axotomy. Neuroscience. 2001;106(1):1619.CrossRefGoogle ScholarPubMed
186.Cummins, TR, Waxman, SG.Downregulation of tetrodotoxin-resistant sodium currents and upregulation of a rapidly repriming tetrodotoxin-sensitive sodium current in small spinal sensory neurons after nerve injury. J Neurosci. 1997;17(10):350314.CrossRefGoogle ScholarPubMed
187.Cummins, TR, Aglieco, F, Renganathan, M, 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. 2001;21(16):595261.CrossRefGoogle Scholar
188.Waxman, SG, Kocsis, JD, Black, JA.Type III sodium channel mRNA is expressed in embryonic but not adult spinal sensory neurons, and is reexpressed following axotomy. J Neurophysiol. 1994;72(1):46670.CrossRefGoogle Scholar
189.Dib-Hajj, S, Black, JA, Felts, P, et al.Down-regulation of transcripts for Na channel α-SNS in spinal sensory neurons following axotomy. PNAS. 1996;93(25):149504.CrossRefGoogle ScholarPubMed
190.Black, JA, Cummins, TR, Plumpton, C, et al.Upregulation of a silent sodium channel after peripheral, but not central, nerve injury in DRG neurons. J Neurophysiol. 1999;82(5):277685.CrossRefGoogle Scholar
191.Dib-Hajj, SD, Fjell, J, Cummins, TR, et al.Plasticity of sodium channel expression in DRG neurons in the chronic constriction injury model of neuropathic pain. Pain. 1999;83(3):591600.CrossRefGoogle ScholarPubMed
192.Kim, CH, Oh, Y, Chung, JM, et al.The changes in expression of three subtypes of TTX sensitive sodium channels in sensory neurons after spinal nerve ligation. Brain Res Mol Brain Res. 2001;95(1-2):15361.CrossRefGoogle ScholarPubMed
193.Dib-Hajj, SD, Black, JA, Cummins, TR, et al.Rescue of alpha-SNS sodium channel expression in small dorsal root ganglion neurons after axotomy by nerve growth factor in vivo. J Neurophysiol. 1998;79(5):266876.CrossRefGoogle ScholarPubMed
194.Boucher, TJ, Okuse, K, Bennett, DL, et al.Potent analgesic effects of GDNF in neuropathic pain states. Science. 2000;290(5489):1247.CrossRefGoogle ScholarPubMed
195.Gold, MS, Weinreich, D, Kim, CS, et al.Redistribution of Na(V)1.8 in uninjured axons enables neuropathic pain. J Neurosci. 2003;23(1):15866.CrossRefGoogle ScholarPubMed
196.Hildebrand, ME, Smith, PL, Bladen, C, et al.A novel slow-inactivation-specific ion channel modulator attenuates neuropathic pain. Pain. 2011;152(4):83343.CrossRefGoogle ScholarPubMed
197.Everill, B, Kocsis, JD.Reduction in potassium currents in identified cutaneous afferent dorsal root ganglion neurons after axotomy. J Neurophysiol. 1999;82(2):7008.CrossRefGoogle ScholarPubMed
198.Kim, DS, Choi, JO, Rim, HD, et al.Downregulation of voltage-gated potassium channel [alpha] gene expression in dorsal root ganglia following chronic constriction injury of the rat sciatic nerve. Mol Brain Res. 2002;105(1-2):14652.CrossRefGoogle ScholarPubMed
199.Sarantopoulos, CD, McCallum, JB, Rigaud, M, et al.Opposing effects of spinal nerve ligation on calcium-activated potassium currents in axotomized and adjacent mammalian primary afferent neurons. Brain Res. 2007;1132(1):8499.CrossRefGoogle ScholarPubMed
200.Boettger, MK, Till, S, Chen, MX, et al.Calcium-activated potassium channel SK1- and IK1-like immunoreactivity in injured human sensory neurones and its regulation by neurotrophic factors. Brain. 2002;125(2):25263.CrossRefGoogle ScholarPubMed
201.Baccei, ML, Kocsis, JD.Voltage-gated calcium currents in axotomized adult rat cutaneous afferent neurons. J Neurophysiol. 2000;83(4):222738.CrossRefGoogle ScholarPubMed
202.Jagodic, MM, Pathirathna, S, Joksovic, PM, et al.Upregulation of the T-type calcium current in small rat sensory neurons after chronic constrictive injury of the sciatic nerve. J Neurophysiol. 2008;99(6):31516.CrossRefGoogle ScholarPubMed
203.Jagodic, MM, Pathirathna, S, Nelson, MT, et al.Cell-specific alterations of T-type calcium current in painful diabetic neuropathy enhance excitability of sensory neurons. J Neurosci. 2007;27(12):330516.CrossRefGoogle ScholarPubMed
204.Titmus, MJ, Faber, DS.Axotomy-induced alterations in the electrophysiological characteristics of neurons. Prog Neurobiol. 1990;35(1):151.CrossRefGoogle ScholarPubMed
205.Charles, P.Mechanisms of analgesia by gabapentin and pregabalin - calcium channel α2-δ[Cavα2-δ] ligands. Pain. 2009;142(1-2):1316.Google Scholar
206.Schmidtko, A, Lötsch, J, Freynhagen, R, et al.Ziconotide for treatment of severe chronic pain. Lancet. 2010;375(9725):156977.CrossRefGoogle ScholarPubMed
207.Abdulla, FA, Smith, PA.Axotomy reduces the effect of analgesic opioids yet increases the effect of nociceptin on dorsal root ganglion neurons. J Neurosci. 1998;18(23):968594.CrossRefGoogle ScholarPubMed
208.Hori, Y, Endo, K, Takahashi, T.Presynaptic inhibitory action of enkephalin on excitatory transmission in superficial dorsal horn of rat spinal cord. J Physiol. 1992;450(1):67385.CrossRefGoogle ScholarPubMed
209.Choe, W, Messinger, RB, Leach, E, et al.TTA-P2 is a potent and selective blocker of T-type calcium channels in rat sensory neurons and a novel antinociceptive agent. Mol Pharmacol. 2011;80(5):90010.CrossRefGoogle Scholar
210.Todorovic, SM, Jevtovic-Todorovic, V.T-type voltage-gated calcium channels as targets for the development of novel pain therapies. Br J Pharmacol. 2011;163(3):48495.CrossRefGoogle ScholarPubMed
211.Yao, H, Donnelly, DF, Ma, C, et al.Upregulation of the hyperpolarization-activated cation current after chronic compression of the dorsal root ganglion. J Neurosci. 2003;23(6):206974.CrossRefGoogle ScholarPubMed
212.Chaplan, SR, Guo, HQ, Lee, DH, et al.Neuronal hyperpolarization-activated pacemaker channels drive neuropathic pain. J Neurosci. 2003;23(4):116978.CrossRefGoogle ScholarPubMed
213.Emery, EC, Young, GT, Berrocoso, EM, et al.HCN2 ion channels play a central role in inflammatory and neuropathic pain. Science. 2011;333(6048):14626.CrossRefGoogle ScholarPubMed
214.Takasu, K, Ono, H, Tanabe, M.Spinal hyperpolarization-activated cyclic nucleotide-gated cation channels at primary afferent terminals contribute to chronic pain. Pain. 2010;151(1):8796.CrossRefGoogle ScholarPubMed
215.McLachlan, EM, Janig, W, Devor, M, et al.Peripheral nerve injury triggers noradrenergic sprouting within dorsal root ganglia. Nature. 1993;363(6429):5436.CrossRefGoogle ScholarPubMed
216.Devor, M, Janig, W, Michaelis, M.Modulation of activity in dorsal root ganglion neurons by sympathetic activation in nerve-injured rats. J Neurophysiol. 1994;71(1):3847.CrossRefGoogle ScholarPubMed
217.Sato, J, Perl, ER.Adrenergic excitation of cutaneous pain receptors induced by peripheral nerve injury. Science. 1991;251(5001):160810.CrossRefGoogle ScholarPubMed
218.Petersen, M, Zhang, J, Zhang, JM, et al.Abnormal spontaneous activity and responses to norepinephrine in dissociated dorsal root ganglion cells after chronic nerve constriction. Pain. 1996;67(2-3):3917.CrossRefGoogle ScholarPubMed
219.Birder, LA, Perl, ER.Expression of α2-adrenergic receptors in rat primary afferent neurones after peripheral nerve injury or inflammation. J Physiol. 1999;515(2):53342.CrossRefGoogle ScholarPubMed
220.Abdulla, FA, Smith, PA.Ectopic alpha2-adrenoceptors couple to N-type Ca2+ channels in axotomized rat sensory neurons. J Neurosci. 1997;17(5):163341.CrossRefGoogle ScholarPubMed
221.Amir, R, Devor, M.Functional cross-excitation between afferent A- and C-neurons in dorsal root ganglia. Neuroscience. 1999;95(1):18995.CrossRefGoogle Scholar
222.Dalal, A, Tata, M, Allegre, G, et al.Spontaneous activity of rat dorsal horn cells in spinal segments of sciatic projection following transection of sciatic nerve or of corresponding dorsal roots. Neuroscience. 1999;94(1):21728.CrossRefGoogle ScholarPubMed
223.Woolf, CJ, Thompson, SW.The induction and maintenance of central sensitization is dependent on N-methyl-D-aspartic acid receptor activation; implications for the treatment of post-injury pain hypersensitivity states. Pain. 1991;44(3):2939.CrossRefGoogle ScholarPubMed
224.Garraway, SM, Petruska, JC, Mendell, LM.BDNF sensitizes the response of lamina II neurons to high threshold primary afferent inputs. Eur J Neurosci. 2003;18(9):246776.CrossRefGoogle ScholarPubMed
225.Baba, H, Ji, RR, Kohno, T, et al.Removal of GABAergic inhibition facilitates polysynaptic A fiber-mediated excitatory transmission to the superficial spinal dorsal horn. Mol Cell Neurosci. 2003;24(3):81830.CrossRefGoogle Scholar
226.Coull, JA, Boudreau, D, Bachand, K, et al.Trans-synaptic shift in anion gradient in spinal lamina I neurons as a mechanism of neuropathic pain. Nature. 2003;424(6951):93842.CrossRefGoogle ScholarPubMed
227.Balasubramanyan, S, Stemkowski, PL, Stebbing, MJ, et al.Sciatic chronic constriction injury produces cell-type-specific changes in the electrophysiological properties of rat substantia gelatinosa neurons. J Neurophysiol. 2006;96(2):57990.CrossRefGoogle ScholarPubMed
228.Sivilotti, L, Woolf, CJ.The contribution of GABAA and glycine receptors to central sensitization: disinhibition and touch-evoked allodynia in the spinal cord. J Neurophysiol. 1994;72(1):16979.CrossRefGoogle ScholarPubMed
229.Costigan, M, Scholz, J, Woolf, CJ.Neuropathic pain: a maladaptive response of the nervous system to damage. Annu Rev Neurosci. 2009;32:132.CrossRefGoogle ScholarPubMed
230.Coull, JA, Beggs, S, Boudreau, D, et al.BDNF from microglia causes the shift in neuronal anion gradient underlying neuropathic pain. Nature. 2005;438(7070):101721.CrossRefGoogle ScholarPubMed
231.Dougherty, KD, Dreyfus, CF, Black, IB.Brain-derived neurotrophic factor in astrocytes, oligodendrocytes, and microglia/macrophages after spinal cord injury. Neurobiol Dis. 2000;7(6 Pt B):57485.CrossRefGoogle ScholarPubMed
232.Lu, VB, Ballanyi, K, Colmers, WF, et al.Neuron type-specific effects of brain-derived neurotrophic factor in rat superficial dorsal horn and their relevance to ‘central sensitization’. J Physiol. 2007;584(Pt 2):54363.CrossRefGoogle ScholarPubMed
233.Lu, VB, Biggs, JE, Stebbing, MJ, et al.Brain-derived neurotrophic factor drives the changes in excitatory synaptic transmission in the rat superficial dorsal horn that follow sciatic nerve injury. J Physiol. 2009;587(Pt 5):101332.CrossRefGoogle ScholarPubMed
234.Prescott, SA, Sejnowski, TJDe Koninck, Y.Reduction of anion reversal potential subverts the inhibitory control of firing rate in spinal lamina I neurons: towards a biophysical basis for neuropathic pain. Mol Pain. 2006;2:32.CrossRefGoogle ScholarPubMed
235.Kerr, BJ, Bradbury, EJ, Bennett, DL, et al.Brain-derived neurotrophic factor modulates nociceptive sensory inputs and NMDA-evoked responses in the rat spinal cord. J Neurosci. 1999;19(12):513848.CrossRefGoogle ScholarPubMed
236.Ulmann, L, Hatcher, JP, Hughes, JP, et al.Up-regulation of P2X4 receptors in spinal microglia after peripheral nerve injury mediates BDNF release and neuropathic pain. J Neurosci. 2008;28(44):112638.CrossRefGoogle ScholarPubMed
237.Scholz, J, Woolf, CJ.The neuropathic pain triad: neurons, immune cells and glia. Nat Neurosci. 2007;10(11):13618.CrossRefGoogle ScholarPubMed
238.Devor, M.Ectopic discharge in A beta afferents as a source of neuropathic pain. Exp Brain Res. 2009;196(1):11528.CrossRefGoogle Scholar
239.Gaudet, A, Popovich, P, Ramer, M.Wallerian degeneration: gaining perspective on inflammatory events after peripheral nerve injury. J Neuroinflammation. 2011;8(1):110.CrossRefGoogle ScholarPubMed
240.Kelley, J.Reactions of neurons to injury. In: Kandel, ER, Schwartz, JH, editors. Principles of neural science. New York: Elsevier Science Publishings Co., Inc. 1985. p. 18795.Google Scholar
241.Ma, C, Shu, Y, Zheng, Z, et al.Similar electrophysiological changes in axotomized and neighboring intact dorsal root ganglion neurons. J Neurophysiol. 2003;89(3):1588602.CrossRefGoogle ScholarPubMed
242.Mosconi, T, Kruger, L.Fixed-diameter polyethylene cuffs applied to the rat sciatic nerve induce a painful neuropathy: ultrastructural morphometric analysis of axonal alterations. Pain. 1996;64(1):3757.CrossRefGoogle Scholar
243.Jones, GJ, Barsby, NL, Cohen, EA, et al.HIV-1 Vpr causes neuronal apoptosis and in vivo neurodegeneration. J Neurosci. 2007;27(14):370311.CrossRefGoogle ScholarPubMed
244.Acharjee, S, Noorbakhsh, F, Stemkowski, PL, et al.HIV-1 viral protein R causes peripheral nervous system injury associated with in vivo neuropathic pain. FASEB J. 2010;24(11):434353.CrossRefGoogle ScholarPubMed
245.Tsao, JW, George, EB, Griffin, JW.Temperature modulation reveals three distinct stages of Wallerian degeneration. J Neurosci. 1999;19(12):471826.CrossRefGoogle ScholarPubMed
246.Luttges, MW, Kelly, PT, Gerren, RA.Degenerative changes in mouse sciatic nerves: electrophoretic and electrophysiologic characterizations. Exp Neurol. 1976;50(3):70633.CrossRefGoogle ScholarPubMed
247.Perrin, FE, Lacroix, S, Aviles-Trigueros, M, et al.Involvement of monocyte chemoattractant protein-1, macrophage inflammatory protein-1alpha and interleukin-1beta in Wallerian degeneration. Brain. 2005;128(Pt 4):85466.CrossRefGoogle ScholarPubMed
248.Wells, MR, Vaidya, U.Morphological alterations in dorsal root ganglion neurons after peripheral axon injury: association with changes in metabolism. Exp Neurol. 1989;104(1):328.CrossRefGoogle ScholarPubMed
249.Sorkin, LS, Schafers, M.Immune cells in peripheral nerve. In: DeLeo, JA, Sorkin, LS, Watkins, LR, editors. Immune and glia regulation of pain. Seattle: IASP Press; 2007. p. 319.Google Scholar
250.Xie, WR, Deng, H, Li, H, et al.Robust increase of cutaneous sensitivity, cytokine production and sympathetic sprouting in rats with localized inflammatory irritation of the spinal ganglia. Neuroscience. 2006;142(3):80922.CrossRefGoogle ScholarPubMed
251.Okuda, T, Ishida, O, Fujimoto, Y, et al.The autotomy relief effect of a silicone tube covering the proximal nerve stump. J Orthop Res. 2006;24(7):142737.CrossRefGoogle ScholarPubMed
252.Zuo, Y, Perkins, NM, Tracey, DJ, et al.Inflammation and hyperalgesia induced by nerve injury in the rat: a key role of mast cells. Pain. 2003;105(3):46779.CrossRefGoogle Scholar
253.Perkins, NM, Tracey, DJ.Hyperalgesia due to nerve injury: role of neutrophils. Neuroscience. 2000;101(3):74557.CrossRefGoogle ScholarPubMed
254.Cui, JG, Holmin, S, Mathiesen, T, et al.Possible role of inflammatory mediators in tactile hypersensitivity in rat models of mononeuropathy. Pain. 2000;88(3):23948.CrossRefGoogle ScholarPubMed
255.Moore, KA, Kohno, T, Karchewski, LA, et al.Partial peripheral nerve injury promotes a selective loss of GABAergic inhibition in the superficial dorsal horn of the spinal cord. J Neurosci. 2002;22(15):672431.CrossRefGoogle ScholarPubMed
256.Chen, Y, Balasubramanyan, S, Lai, AY, et al.Effects of sciatic nerve axotomy on excitatory synaptic transmission in rat substantia gelatinosa. J Neurophysiol. 2009;102(6):320315.CrossRefGoogle ScholarPubMed
257.Tominaga, M, Caterina, MJ, Malmberg, AB, et al.The cloned capsaicin receptor integrates multiple pain-producing stimuli. Neuron. 1998;21(3):53143.CrossRefGoogle ScholarPubMed
258.Chen, CC, Akopian, AN, Sivilotti, L, et al.A P2X purinoceptor expressed by a subset of sensory neurons. Nature. 1995;377(6548):42831.CrossRefGoogle ScholarPubMed
259.Chen, CC, England, S, Akopian, AN, et al.A sensory neuron-specific, proton-gated ion channel. PNAS. 1998;95(17):102405.CrossRefGoogle ScholarPubMed
260.Nakamura-Craig, M, Gill, BK.Effect of neurokinin A, substance P and calcitonin gene related peptide in peripheral hyperalgesia in the rat paw. Neurosci Lett. 1991;124(1):4951.CrossRefGoogle ScholarPubMed
261.Carlton, SM, Zhou, S, Coggeshall, RE.Localization and activation of substance P receptors in unmyelinated axons of rat glabrous skin. Brain Res. 1996;734(1-2):1038.CrossRefGoogle ScholarPubMed
262.Dray, A, Pinnock, RD.Effects of substance P on adult rat sensory ganglion neurones in vitro. Neurosci Lett. 1982;33(1):616.CrossRefGoogle ScholarPubMed
263.Abdulla, FA, Stebbing, MJ, Smith, PA.Effects of substance P on excitability and ionic currents of normal and axotomized rat dorsal root ganglion neurons. Eur J Neurosci. 2001;13(3):54552.CrossRefGoogle ScholarPubMed
264.Li, HS, Zhao, ZQ.Small sensory neurons in the rat dorsal root ganglia express functional NK-1 tachykinin receptor. Eur J Neurosci. 1998;10(4):12929.CrossRefGoogle ScholarPubMed
265.Yonehara, N, Yoshimura, M.Influence of painful chronic neuropathy on neurogenic inflammation. Pain. 2001;92(1-2):25965.CrossRefGoogle ScholarPubMed
266.Cahill, CM, Coderre, TJ.Attenuation of hyperalgesia in a rat model of neuropathic pain after intrathecal pre- or post-treatment with a neurokinin-1 antagonist. Pain. 2002;95(3):27785.CrossRefGoogle ScholarPubMed
267.Lee, SE, Kim, JH.Involvement of substance P and calcitonin gene-related peptide in development and maintenance of neuropathic pain from spinal nerve injury model of rat. Neurosci Res. 2007;58(3):2459.CrossRefGoogle ScholarPubMed
268.Jang, JH, Nam, TS, Paik, KS, et al.Involvement of peripherally released substance P and calcitonin gene-related peptide in mediating mechanical hyperalgesia in a traumatic neuropathy model of the rat. Neurosci Lett. 2004;360(3):12932.CrossRefGoogle Scholar
269.Herzberg, U, Eliav, E, Dorsey, JM, et al.NGF involvement in pain induced by chronic constriction injury of the rat sciatic nerve. Neuroreport. 1997;8(7):161318.CrossRefGoogle ScholarPubMed
270.Zhou, XF, Deng, YS, Chie, E, et al.Satellite-cell-derived nerve growth factor and neurotrophin-3 are involved in noradrenergic sprouting in the dorsal root ganglia following peripheral nerve injury in the rat. Eur J Neurosci. 1999;11(5):171122.CrossRefGoogle ScholarPubMed
271.Ro, LS, Chen, ST, Tang, LM, et al.Effect of NGF and anti-NGF on neuropathic pain in rats following chronic constriction injury of the sciatic nerve. Pain. 1999;79(2-3):26574.CrossRefGoogle ScholarPubMed
272.Chuang, HH, Prescott, ED, Kong, H, et al.Bradykinin and nerve growth factor release the capsaicin receptor from PtdIns(4,5)P2-mediated inhibition. Nature. 2001;411(6840):95762.CrossRefGoogle ScholarPubMed
273.Shu, X, Mendell, LM.Acute sensitization by NGF of the response of small-diameter sensory neurons to capsaicin. J Neurophysiol. 2001;86(6):29318.CrossRefGoogle ScholarPubMed
274.Ji, RR, Samad, TA, Jin, SX, et al.p38 MAPK activation by NGF in primary sensory neurons after inflammation increases TRPV1 levels and maintains heat hyperalgesia. Neuron. 2002;36(1):5768.CrossRefGoogle ScholarPubMed
275.Zhang, YH, Nicol, GD.NGF-mediated sensitization of the excitability of rat sensory neurons is prevented by a blocking antibody to the p75 neurotrophin receptor. Neurosci Lett. 2004;366(2):18792.CrossRefGoogle Scholar
276.Zhang, YH, Vasko, MR, Nicol, GD.Ceramide, a putative second messenger for nerve growth factor, modulates the TTX-resistant Na+ current and delayed rectifier K+ current in rat sensory neurons. J Physiol. 2002;544(2):385402.CrossRefGoogle ScholarPubMed
277.Amadesi, S, Nie, J, Vergnolle, N, et al.Protease-activated receptor 2 sensitizes the capsaicin receptor transient receptor potential vanilloid receptor 1 to induce hyperalgesia. J Neurosci. 2004;24(18):430012.CrossRefGoogle ScholarPubMed
278.Linley, JE, Rose, K, Patil, M, et al.Inhibition of M current in sensory neurons by exogenous proteases: a signaling pathway mediating inflammatory nociception. J Neurosci. 2008;28(44):112409.CrossRefGoogle ScholarPubMed
279.Alier, KA, Endicott, JA, Stemkowski, PL, et al.Intrathecal administration of proteinase-activated receptor-2 agonists produces hyperalgesia by exciting the cell bodies of primary sensory neurons. J Pharmacol Exp Ther. 2008;324(1):22433.CrossRefGoogle ScholarPubMed
280.Kress, M, Guenther, S.Role of [Ca2+]i in the ATP-induced heat sensitization process of rat nociceptive neurons. J Neurophysiol. 1999;81(6):261219.CrossRefGoogle ScholarPubMed
281.Tominaga, M, Wada, M, Masu, M.Potentiation of capsaicin receptor activity by metabotropic ATP receptors as a possible mechanism for ATP-evoked pain and hyperalgesia. PNAS. 2001;98(12):69516.CrossRefGoogle ScholarPubMed
282.Cesare, P, McNaughton, P.A novel heat-activated current in nociceptive neurons and its sensitization by bradykinin. PNAS. 1996;93(26):154359.CrossRefGoogle ScholarPubMed
283.Dai, Y, Wang, S, Tominaga, M, et al.Sensitization of TRPA1 by PAR2 contributes to the sensation of inflammatory pain. J Clin Invest. 2007;117(7):197987.CrossRefGoogle Scholar
284.Cenac, N, Andrews, CN, Holzhausen, M, et al.Role for protease activity in visceral pain in irritable bowel syndrome. J Clin Invest. 2007;117(3):63647.CrossRefGoogle ScholarPubMed
285.Salmon, JA, Higgs, GA.Prostaglandins and leukotrienes as inflammatory mediators. Br Med Bull. 1987;43(2):28596.CrossRefGoogle ScholarPubMed
286.Kiefer, W, Dannhardt, G.COX-2 inhibition and pain management: a review summary. Expert Rev Clin Immunol. 2005;1(3):43142.CrossRefGoogle ScholarPubMed
287.Ferreira, SH, Nakamura, M, de Abreu Castro, MS.The hyperalgesic effects of prostacyclin and prostaglandin E2. Prostaglandins. 1978;16(1):317.CrossRefGoogle ScholarPubMed
288.Higgs, EA, Moncada, S, Vane, JR.Inflammatory effects of prostacyclin (PGI2) and 6-oxo-PGF1[alpha] in the rat paw. Prostaglandins. 1978;16(2):15362.CrossRefGoogle Scholar
289.Schepelmann, K, Linger, K, Schaible, HG, et al.Inflammatory mediators and nociception in the joint: Excitation and sensitization of slowly conducting afferent fibers of cat’s knee by prostaglandin I2. Neuroscience. 1992;50(1):23747.CrossRefGoogle ScholarPubMed
290.Birrell, GJ, McQueen, DS, Iggo, A, et al.PGI2-induced activation and sensitization of articular mechanonociceptors. Neurosci Lett. 1991;124(1):58.CrossRefGoogle ScholarPubMed
291.Devor, M, White, DM, Goetzl, EJ, et al.Eicosanoids, but not tachykinins, excite C-fiber endings in rat sciatic nerve-end neuromas. Neuroreport. 1992;3(1):214.CrossRefGoogle Scholar
292.Oida, H, Namba, T, Sugimoto, Y, et al.In situ hybridization studies of prostacyclin receptor mRNA expression in various mouse organs. Br J Pharmacol. 1995;116(7):282837.CrossRefGoogle ScholarPubMed
293.Sugimoto, Y, Shigemoto, R, Namba, T, et al.Distribution of the messenger RNA for the prostaglandin e receptor subtype ep3 in the mouse nervous system. Neuroscience. 1994;62(3):91928.CrossRefGoogle ScholarPubMed
294.Matsumura, K, Watanabe, Y, Onoe, H, et al.Prostacyclin receptor in the brain and central terminals of the primary sensory neurons: an autoradiographic study using a stable prostacyclin analogue [3H]iloprost. Neuroscience. 1995;65(2):493503.CrossRefGoogle Scholar
295.Schaible, HG, Schmidt, RF.Excitation and sensitization of fine articular afferents from cat’s knee joint by prostaglandin E2. J Physiol. 1988;403:91104.CrossRefGoogle ScholarPubMed
296.Nicol, GD, Cui, M.Enhancement by prostaglandin E2 of bradykinin activation of embryonic rat sensory neurones. J Physiol. 1994;480(Pt 3):48592.CrossRefGoogle ScholarPubMed
297.Nicol, GD, Vasko, MR, Evans, AR.Prostaglandins suppress an outward potassium current in embryonic rat sensory neurons. J Neurophysiol. 1997;77(1):16776.CrossRefGoogle ScholarPubMed
298.Devor, M, Govrin-Lippmann, R, Angelides, K.Na+ channel immunolocalization in peripheral mammalian axons and changes following nerve injury and neuroma formation. J Neurosci. 1993;13(5):197692.CrossRefGoogle ScholarPubMed
299.Gold, MS, Reichling, DB, Shuster, MJ, et al.Hyperalgesic agents increase a tetrodotoxin-resistant Na+ current in nociceptors. PNAS. 1996;93(3):110812.CrossRefGoogle ScholarPubMed
300.Gold, MS, Levine, JD.DAMGO inhibits prostaglandin E2-induced potentiation of a TTX-resistant Na+ current in rat sensory neurons in vitro. Neurosci Lett. 1996;212(2):836.CrossRefGoogle ScholarPubMed
301.Khasar, SG, Gold, MS, Levine, JD.A tetrodotoxin-resistant sodium current mediates inflammatory pain in the rat. Neurosci Lett. 1998;256(1):1720.CrossRefGoogle ScholarPubMed
302.Ingram, SL, Williams, JT.Modulation of the hyperpolarization-activated current (Ih) by cyclic nucleotides in guinea-pig primary afferent neurons. J Physiol. 1996;492(Pt 1):97106.CrossRefGoogle ScholarPubMed
303.Moriyama, T, Higashi, T, Togashi, K, et al.Sensitization of TRPV1 by EP1 and IP reveals peripheral nociceptive mechanism of prostaglandins. Mol Pain. 2005;1(1):3.CrossRefGoogle ScholarPubMed
304.Taiwo, YO, Levine, JD.Characterization of the arachidonic acid metabolites mediating bradykinin and noradrenaline hyperalgesia. Brain Res. 1988;458(2):4026.CrossRefGoogle ScholarPubMed
305.Sommer, C, Kress, M.Recent findings on how proinflammatory cytokines cause pain: peripheral mechanisms in inflammatory and neuropathic hyperalgesia. Neurosci Lett. 2004;361(1-3):1847.CrossRefGoogle ScholarPubMed
306.Leung, L, Cahill, CM.TNF-alpha and neuropathic pain-a review. J Neuroinflammation. 2010;7:27.CrossRefGoogle ScholarPubMed
307.Copray, JC, Mantingh, I, Brouwer, N, et al.Expression of interleukin-1 beta in rat dorsal root ganglia. J Neuroimmunol. 2001;118(2):20311.CrossRefGoogle ScholarPubMed
308.Liu, L, Yang, TM, Liedtke, W, et al.Chronic IL-1beta signaling potentiates voltage-dependent sodium currents in trigeminal nociceptive neurons. J Neurophysiol. 2006;95(3):147890.CrossRefGoogle ScholarPubMed
309.Binshtok, AM, Wang, H, Zimmermann, K, et al.Nociceptors are interleukin-1beta sensors. J Neurosci. 2008;28(52):1406273.CrossRefGoogle ScholarPubMed
310.Takeda, M, Tanimoto, T, Kadoi, J, et al.Enhanced excitability of nociceptive trigeminal ganglion neurons by satellite glial cytokine following peripheral inflammation. Pain. 2007;129(1-2):15566.CrossRefGoogle Scholar
311.Nadeau, S, Filali, M, Zhang, J, et al.Functional recovery after peripheral nerve injury is dependent on the pro-inflammatory cytokines IL-1β and TNF: implications for neuropathic pain. J Neurosci. 2011;31(35):1253342.CrossRefGoogle ScholarPubMed
312.Rotshenker, S, Aamar, S, Barak, V.Interleukin-1 activity in lesioned peripheral nerve. J Neuroimmunol. 1992;39(1-2):7580.CrossRefGoogle ScholarPubMed
313.Kawasaki, Y, Xu, ZZ, Wang, X, et al.Distinct roles of matrix metalloproteases in the early- and late-phase development of neuropathic pain. Nat Med. 2008;14(3):3316.CrossRefGoogle ScholarPubMed
314.Stemkowski, PL, Smith, PA.Long-term IL-1β exposure causes subpopulation dependent alterations in rat dorsal root ganglion neuron excitability. J Neurophysiol. 2012;107:158697.CrossRefGoogle ScholarPubMed
315.Temporin, K, Tanaka, H, Kuroda, Y, et al.Interleukin-1 beta promotes sensory nerve regeneration after sciatic nerve injury. Neurosci Lett. 2008;440(2):1303.CrossRefGoogle ScholarPubMed
316.Temporin, K, Tanaka, H, Kuroda, Y, et al.IL-1[beta] promotes neurite outgrowth by deactivating RhoA via p38 MAPK pathway. Biochem Biophys Res Commun. 2008;365(2):37580.CrossRefGoogle ScholarPubMed
317.Horie, H, Sakai, I, Akahori, Y, et al.IL-1 beta enhances neurite regeneration from transected-nerve terminals of adult rat DRG. Neuroreport. 1997;8(8):19559.CrossRefGoogle ScholarPubMed
318.Sommer, C, Myers, RR.Neurotransmitters in the spinal cord dorsal horn in a model of painful neuropathy and in nerve crush. Acta Neuropathol. 1995;90(5):47885.CrossRefGoogle Scholar
319.Marchand, JE, Wurm, WH, Kato, T, et al.Altered tachykinin expression by dorsal root ganglion neurons in a rat model of neuropathic pain. Pain. 1994;58(2):21931.CrossRefGoogle Scholar
320.Bennett, AD, Chastain, KM, Hulsebosch, CE.Alleviation of mechanical and thermal allodynia by CGRP(8-37) in a rodent model of chronic central pain. Pain. 2000;86(1-2):16375.CrossRefGoogle Scholar
321.Baranowski, AP, Priestley, JV, McMahon, S.Substance P in cutaneous primary sensory neurons-a comparison of models of nerve injury that allow varying degrees of regeneration. Neuroscience. 1993;55(4):102536.CrossRefGoogle ScholarPubMed
322.Noguchi, K, Dubner, R, De Leon, M, et al.Axotomy induces preprotachykinin gene expression in a subpopulation of dorsal root ganglion neurons. J Neurosci Res. 1994;37(5):596603.CrossRefGoogle Scholar
323.Noguchi, K, Kawai, Y, Fukuoka, T, et al.Substance P induced by peripheral nerve injury in primary afferent sensory neurons and its effect on dorsal column nucleus neurons. J Neurosci. 1995;15(11):763343.CrossRefGoogle ScholarPubMed
324.Michael, GJ, Averill, S, Nitkunan, A, et al.Nerve growth factor treatment increases brain-derived neurotrophic factor selectively in TrkA-expressing dorsal root ganglion cells and in their central terminations within the spinal cord. J Neurosci. 1997;17(21):847690.CrossRefGoogle ScholarPubMed
325.Tonra, JR, Curtis, R, Wong, V, et al.Axotomy upregulates the anterograde transport and expression of brain-derived neurotrophic factor by sensory neurons. J Neurosci. 1998;18(11):437483.CrossRefGoogle ScholarPubMed