The role of slow and persistent TTX-resistant sodium currents in acute tumor necrosis factor-α-mediated increase in nociceptors excitability

Sagi Gudes, Omer Barkai, Yaki Caspi, Ben Katz, Shaya Lev, Alexander M Binshtok, Sagi Gudes, Omer Barkai, Yaki Caspi, Ben Katz, Shaya Lev, Alexander M Binshtok

Abstract

Tetrodotoxin-resistant (TTX-r) sodium channels are key players in determining the input-output properties of peripheral nociceptive neurons. Changes in gating kinetics or in expression levels of these channels by proinflammatory mediators are likely to cause the hyperexcitability of nociceptive neurons and pain hypersensitivity observed during inflammation. Proinflammatory mediator, tumor necrosis factor-α (TNF-α), is secreted during inflammation and is associated with the early onset, as well as long-lasting, inflammation-mediated increase in excitability of peripheral nociceptive neurons. Here we studied the underlying mechanisms of the rapid component of TNF-α-mediated nociceptive hyperexcitability and acute pain hypersensitivity. We showed that TNF-α leads to rapid onset, cyclooxygenase-independent pain hypersensitivity in adult rats. Furthermore, TNF-α rapidly and substantially increases nociceptive excitability in vitro, by decreasing action potential threshold, increasing neuronal gain and decreasing accommodation. We extended on previous studies entailing p38 MAPK-dependent increase in TTX-r sodium currents by showing that TNF-α via p38 MAPK leads to increased availability of TTX-r sodium channels by partial relief of voltage dependence of their slow inactivation, thereby contributing to increase in neuronal gain. Moreover, we showed that TNF-α also in a p38 MAPK-dependent manner increases persistent TTX-r current by shifting the voltage dependence of activation to a hyperpolarized direction, thus producing an increase in inward current at functionally critical subthreshold voltages. Our results suggest that rapid modulation of the gating of TTX-r sodium channels plays a major role in the mediated nociceptive hyperexcitability of TNF-α during acute inflammation and may lead to development of effective treatments for inflammatory pain, without modulating the inflammation-induced healing processes.

Keywords: DRG; inflammatory pain; nociceptor; sodium (Na+) current; tumor necrosis factor.

Copyright © 2015 the American Physiological Society.

Figures

Fig. 1.
Fig. 1.
Tumor necrosis factor (TNF)-α produces acute cyclooxygenase (COX)-2-independent hypersensitivity to noxious mechanical and thermal stimuli. Latency for paw withdrawal from hot plate (52°C), normalized to baseline (A) and mechanical threshold for paw withdrawal, normalized to baseline, measured by von Frey filaments (B), after intraplantar injection of vehicle, followed, 20 min later, by intraplantar injection of saline (Vehicle-Saline group, black squares); vehicle followed, 20 min later, by intraplantar injection of TNF-α (Vehicle-TNF-α group open circles); COX antagonist, ibuprofen, followed, 20 min later, by intraplantar injection of saline (Ibuprofen-Saline group, gray triangles); or ibuprofen followed by intraplantar injection of TNF-α (Ibuprofen-TNF-α group, light gray inverted triangles). The values are means ± SE; n = 6 for each group. Time point “0” refers to the baseline. Black asterisks, comparison between Vehicle-TNF-α group and Vehicle-Saline group; gray asterisks, comparison between Ibuprofen-TNF-α group and Ibuprofen-Saline group. Nonsignificant (ns), P > 0.05; ***P < 0.001; **P < 0.01; *P < 0.05; two-way ANOVA followed by Bonferroni post-test.
Fig. 2.
Fig. 2.
Application of TNF-α increases the excitability of nociceptor-like dorsal root ganglion (DRG) neurons. A: representative traces of membrane voltage responses to current steps, recorded from the same small nociceptor-like DRG neurons, before and 10 min after bath application of TNF-α. Note that, following the application of TNF-α, subthreshold stimuli become suprathreshold (Rheobasebefore = 187.3 ± 61.6 pA; RheobaseTNF-α = 139.1 ± 52.8 pA, P < 0.01, n = 11). B: frequency-intensity (f–I) curves calculated from the same neurons, before (gray squares) and 10 min after application of TNF-α (black circles), expressed as spike frequency (means ± SE) evoked by suprathreshold stimulus and plotted against stimuli intensity normalized to rheobase of each analyzed neuron. Note the significantly increased spike frequency (***P < 0.001, two-way ANOVA, n = 10) and increase in slope (m) of the f–I curve following application of TNF-α (P < 0.001, paired t-test, n = 11). C: representative traces of the membrane voltage responses to 1.5-s depolarizing current ramps (333 pA/s) recorded from the same neuron before and after application of TNF-α. Note that TNF-α reduced first spike latency (FSL) and increased action potential (AP) frequency. D: histogram plotting the spike count before and 10 min after application of TNF-α measured from traces as in C. Note the increase in the number of spikes (means ± SE, **P < 0.01, paired t-test, n = 10). E: AP thresholds for generation of APs, obtained by the phase plot analysis of AP, before and 10 min after application of TNF-α (means ± SE, **P < 0.01, paired t-test, n = 11).
Fig. 3.
Fig. 3.
TNF-α increases the amplitude of tetrodotoxin-resistant (TTX-r) slow and TTX-r persistent sodium currents in nociceptor-like DRG neurons. A–C, left: representative traces of total (A), persistent TTX-r (B) and slow TTX-r (C) sodium currents recorded from the same neuron before (gray traces) and 1 min after application of TNF-α (black traces). The currents were elicited by protocols represented in the inset beneath the traces (for the detailed protocols, see materials and methods). The dashed line indicates the zero current level. Currents are adjusted according to the baseline before the activation step. Traces (150 μs) were blanked to remove residual uncompensated capacitance transient currents. Note the different scale bars between A, B and C. Right: ratio of the peak of total (A), persistent TTX-r (B) and slow TTX-r (C) sodium currents after TNF-α application (Ipeak,TNF-α) to the peak current before (Ipeak,before) from all measured cells (means ± SE, nTotal = 14, nTTX-r per = 5, nTTX-r = 30, paired t-test, ***P < 0.001). Dotted lines indicate the level of Ipeak,before. D: time course of changes of the peak amplitude of TTX-r sodium current assessed by 150-ms depolarizing steps from −70 mV to 0 mV during a 20-min application of TNF-α (black circles) or vehicle (gray squares) (n = 11 for time points −2.5 to 17.5; n = 7 for time point 20; n = 5 for time points 25 to 30; means ± SE). Time point −2.5 refers to data collected shortly after establishing the whole cell configuration when vehicle was applied. Time points 0 to 20 indicate time of application of TNF-α. Note fast onset of the effect of TNF-α (1 min) and decrease in sodium current amplitude when recorded without TNF-α (n = 9 for time points −2.5 to 10; n = 6 for time point 12.5 to 30; n = 5 for time points 25 to 30; means ± SE). Peak current normalized to peak current amplitude detected at −2.5 time point (3.1 ± 1.1 nA for the TNF-α group, 3.2 ± 0.6 nA, for the vehicle group, **P < 0.01, two-way ANOVA, n = 30) is marked by dashed line.
Fig. 4.
Fig. 4.
Application of TNF-α reduces the threshold and increases the maximal upstroke velocity of the AP, in the presence of TTX. A: representative phase plots of rate of change of the membrane potential (dV/dt) during an AP vs. membrane potential (Vm) recorded from the same neuron before (gray) and 10 min after application of TNF-α (black), in the presence of TTX. Note the leftward shift of the threshold voltage (gray arrows, quantified in B) and increase in maximal velocity, (dV/dt)max, of the upstroke of the AP (black arrows, quantified in C). Inset, traces of APs, before (gray) and after (black) application of TNF-α, which have been used for the generation of the phase plot, presented in A. The arrows indicate the estimated counterpart point presented in the phase plot graph. B: the thresholds for generation of APs, in the presence of TTX, calculated from the same neurons before and 10 min after application of TNF-α (means ± SE, **P < 0.01, paired t-test, n = 8). C: the maximal velocity (dV/dt)max of the upstroke of the AP before (gray) and after (black) application of TNF-α (means ± SE, **P < 0.01, paired t-test, n = 8).
Fig. 5.
Fig. 5.
TNF-α shifts the activation of TTX-r persistent in hyperpolarized direction. A: averaged peak current-voltage (I–V) characteristics of TTX-r persistent sodium current, before (open squares) and 1 min after application of TNF-α (black circles). To estimate the change in the voltage dependence of activation, we calculated the voltage shift, which entails the minimum “distance” between the sum of control current values [Ibefore (V)] and their approximate equivalents after TNF-α [ITNF-α (V + vstep)]. To this end, Ibefore (V) values were shifted in steps with 0.01-mV increments towards ITNF-α (V), while ITNF-α (V) was kept constant. The shift (arrow) of 10 mV gave the best fit (see materials and methods). Inset: representative traces of TTX-r persistent sodium currents, recorded in the presence of TTX and a blocker of voltage-gated sodium (Nav) 1.8 channels, A803467, elicited by 120-ms steps from a holding potential of −90 mV to −50 mV (light gray), −40 mV (gray) and −30 mV (black) before and 1 min after application of TNF-α. Currents are adjusted according to the baseline before the activation step. The dashed line indicates the zero current level. B: representative instantaneous I–V curve during a slow (36 mV/s) depolarizing ramp from −70 to +20 mV before (gray) and 1 min after application of TNF-α (black). Note “W” shaped inward current with clear early and late components. TNF-α increased the amplitude of both early and late components of the inward current (C) (**P < 0.01, *P < 0.05, paired t-test, n = 6) and shifted the onset (circles) and peak activation (squares) of early, but not the late, component to more negative potentials (D) (***P < 0.001, **P < 0.01, *P > 0.05, paired t-test, n = 6).
Fig. 6.
Fig. 6.
TNF-α increases TTX-r sodium currents in nociceptor-like DRG neurons by modulating voltage dependence of slow inactivation. A: representative families of TTX-r sodium currents evoked by a series of 150-ms, 10-mV depolarizing voltage steps from a holding potential of −70 mV (top) or −120 mV (bottom), recorded from the same neurons before and 1 min after application of TNF-α. The dashed lines indicate the zero current level. The dotted lines indicate the peak current level before application of TNF-α. B: histogram plotting the ratio of the peak sodium currents 1 min after TNF-α application (Ipeak,TNF-α) to the peak current before application (Ipeak,before) when elicited from a holding potential (Vholding) of −70 mV or −120 mV (means ± SE, ***P < 0.001, **P < 0.01, Student's t-test, n = 30). For A and B, currents are adjusted according to the baseline before the activation step. Note, significant increase in current at holding of −70 mV and significant decrease in current at holding of −120 mV. C: activation [conductance/maximum conductance (G/Gmax)] and availability [current/maximal current (I/Imax)] curves show no change in voltage dependence of activation (P = 0.607, paired t-test between the values of V1/2,before and V1/2,TNF-α, n = 30) and fast inactivation (P = 0.718, paired t-test between the values of V1/2,before and V1/2,TNF-α, n = 14), respectively, following TNF-α. D: superimposed representative traces of TTX-r sodium currents before (open squares) and 3 min after application of TNF-α (closed circles). Currents were elicited by 10-ms test steps to 0 mV, after 5-s conditioning pulses (Vcond) to the indicated voltages. Note that the substantial current reduction, following depolarized Vcond, was diminished by application of TNF-α. Currents are adjusted according to the baseline before the activation step. The dashed line indicates the zero current level. Inset: stimulus protocol (see materials and methods). The values of the current following the test step (Itest) were divided by the maximal current (Imax) and plotted as means ± SE against Vcond to calculate the fraction of channels available for activation at each holding voltage (E) (n = 25). The lines are the best fits to the Boltzmann function (see materials and methods). Note the increase in fraction of available channels following TNF-α [refer to Table 2 for values of the midpoints of voltage dependence (V1/2), values of slope and statistical analysis].
Fig. 7.
Fig. 7.
In compartmental simulations, rightward shift in voltage dependence of slow inactivation, equivalent to one produced by TNF-α, leads to increase in AP firing rate. A: a Markovian state model for TTX-r sodium channel represents two parallel planes: the first, a classic Hodgkin-Huxley like kinetic scheme with fast voltage-dependent transition; and the second, describing the slow inactivation states with much slower kinetics (see materials and methods). O denotes the open state; C denotes the closed states; IC and SC represent the fast and slow inactivation states, respectively; and SIC is the state in which the channel is inactivated by both slow and fast inactivation mechanisms. Activation occurs with the independent movement of gates, with voltage-dependent transitions between the multiple closed states. B: responses of simulated neuron to 2,000-ms hyperpolarizing and depolarizing suprathreshold current steps without (left) and with (right) 3.32-mV rightward shift in the voltage dependence of slow inactivation of slow Nav 1.8 channels (see materials and methods). A 3.32-mV shift was used to match the shift in voltage dependence in slow inactivation produced by TNF-α (Fig. 6 and Table 2). Note the increase in AP firing following the same stimuli (shown below) when rightward shift is applied. C: f–I curves calculated with (black circles) and without (gray squares) a 3.32-mV rightward shift in the voltage dependence of slow inactivation of Nav 1.8 channels, expressed as spike frequency evoked by suprathreshold stimulus and plotted against stimuli intensity normalized to rheobase. Note the significantly increased spike frequency and increase in slope (m) of the f–I curve following the shift. D: rightward shift of 3.32 mV in voltage dependence of slow inactivation in simulated neuron produced increased firing rate in response to 1.5-s-long depolarizing current ramps (333 pA/s, shown below). E and F: histograms plotting the spike count (E) and the maximal velocity (dV/dt)max of the upstroke of the AP (F) with (black) and without (gray) 3.32-mV depolarizing shift in the voltage dependence of slow inactivation of Nav 1.8 channels.
Fig. 8.
Fig. 8.
TNF-α-mediated relief of voltage dependence of slow inactivation and increase in nociceptive excitability requires p38 MAPK. A: voltage dependence of slow inactivation of TTX-r sodium currents, expressed as peaks of TTX-r sodium currents during a step to 0 mV, normalized to Imax (fraction available, means ± SE) and plotted against conditioning voltages, before (gray squares) and 2 min after application of p38 MAPK antagonist SB202190, followed by a 1-min coapplication with TNF-α (open circles, paired t-test between the values of V1/2,before and V1/2,TNF-α, P > 0.05, n = 14, see also Table 2). Inset: representative traces of TTX-r sodium currents, recorded in the presence of TTX, elicited by 150-ms steps from a holding potential of −70 mV to 0 mV before (gray) and 2 min after application of p38 MAPK antagonist SB202190, followed by a 1 min coapplication with TNF-α (black). Currents are adjusted according to the baseline before the activation step. The dashed line indicates the zero current level. The dotted line indicates the peak current level before application of TNF-α. B: voltage dependence of slow inactivation of TTX-r sodium currents, assessed as explained in A, before (gray squares) and 2 min after application of the inactive analog, SB202474, followed by a 1-min coapplication with TNF-α (black, paired t-test between the values of V1/2,before and V1/2,TNF-α, P < 0.0001, n = 18; see also Table 2). Inset: representative traces of TTX-r sodium currents, recorded in the presence of 300 nM TTX, elicited by 150-ms steps from a holding potential of −70 mV to 0 mV before (gray) and 2 min after application of the inactive analog SB202424, followed by 1-min coapplication with TNF-α (black). The dashed line indicates the zero current level. The dotted line indicates the peak current level before application of TNF-α. C: the ratio of peak TTX-r sodium current before application of treatment (INa,before) to the peak current after application (INa,treatment) of SB202190 + TNF-α (n = 14) and SB202474 + TNF-α (n = 18). The values are means ± SE; ns, P > 0.05, ***P < 0.001, one-sample t-test comparing the values before and after the application of treatment. D: representative traces of current clamp recordings from the same nociceptor-like DRG neuron, following 500 pA, 1.5-s depolarizing ramps before (left) and 3 min after application of p38 MAPK antagonist SB202190, followed by a 10-min coapplication with TNF-α (right). FSL, first spike latency. E: SB202190 prevents the effect of TNF-α on spike count (means ± SE, **P < 0.01, paired t-test, n = 8) and threshold for generation of AP (F) (means ± SE, ns, P > 0.05, paired t-test, n = 8). G: latency for paw withdrawal from hot plate (5°C), normalized to baseline, after intraplantar injection of vehicle followed, 20 min later, by intraplantar injection of saline (Vehicle-Saline, black squares); intraplantar injection of vehicle followed, 20 min later, by intraplantar injection of TNF-α (Vehicle-TNF-α, open circles); intraplantar injection of p38 MAPK inhibitor, SB203580, followed, 20 min later, by intraplantar injection of saline (SB203580-Saline, gray triangles); or intraplantar injection of SB203580 followed by intraplantar injection of TNF-α (SB203580-TNF-α, light gray inverted triangles). The values are means ± SE. Time point “0” refers to the baseline. Gray asterisks, comparison between SB203580-Saline group to SB203580-TNF-α group. ns, P > 0.05; **P < 0.01; *P < 0.05; two-way ANOVA followed by Bonferroni post-test, n = 6 for each group. H: mechanical threshold for paw withdrawal, normalized to baseline, obtained by von Frey filaments. Time point “0” refers to the baseline. Symbols are as in G.
Fig. 9.
Fig. 9.
p38 MAPK modulates voltage dependence of slow inactivation. A: voltage dependence of slow inactivation of TTX-r sodium currents, expressed as peaks of TTX-r sodium currents during a step to 0 mV, normalized to Imax (fraction available, means ± SE) and plotted against conditioning voltages before (gray squares) and 3 min after application of SB202190 (open circles, paired t-test, P < 0.05, n = 7. See Table 2 for the values of the V1/2 and slope and statistical analysis). Inset: representative families of TTX-r sodium currents, recorded in the presence of TTX, elicited by 150-ms, 10-mV steps from a holding potential of −70 mV before (gray) and 3 min after application of SB202190 (black). Currents are adjusted according to the baseline before the activation step. The dashed line indicates the zero current level. The dotted line indicates the peak current level before application of SB202190. B: the ratio of peak TTX-r sodium current before application of treatment (INa,before) to the peak current after application (INa,treatment) of vehicle (n = 11) or SB202190 alone (n = 6). Values are means ± SE; **P < 0.01, one sample t-test comparing the values before and after the application of treatment. C: voltage dependence of slow inactivation of TTX-r sodium currents, expressed as peaks of TTX-r sodium currents during steps to 0 mV, normalized to Imax (fraction available, means ± SE) and plotted against conditioning voltages, measured 30 min after pretreatment with the p38 MAPK activator anisomycin (open circles, n = 26) or vehicle (black squares, n = 20), P < 0.01, two-sampled t-test; see also Table 3 for numerical data. Inset: histograms of the distribution of the values of V1/2 of voltage dependence of slow inactivation of TTX-r sodium currents, recorded form neurons preincubated for 30 min with vehicle (gray, n = 20) or 10 μg/ml anisomycin (white, n = 26). Fitting of the distribution with the amplitude version of Gaussian peak function, y = y0 + A × exp(−0.5 × {[V1/2 − V1/2(c)]/w})2, where A is the peak number of cells and V1/2(c) is the value of V1/2 which corresponds to A; black solid line for vehicle; gray solid line for anisomycin, reveals that anisomycin produced substantial rightward shift of population of V1/2 values [V1/2(c),Vehicle = −30.3 ± 1.5 mV; V1/2(c),Anisomycin = −21.2 ± 0.9 mV]. D: the peak TTX-r current density, measured from currents elicited by depolarizing steps to 0 mV from a holding potential of −70 mV, in cells pretreated with anisomycin or vehicle. Values are means ± SE, *P < 0.01, two sampled t-test, n = 26 for anisomycin group, n = 20 for vehicle group.
Fig. 10.
Fig. 10.
Pharmacological enhancement of voltage dependence of slow inactivation reverses the TNF-α-mediated relief of slow inactivation and diminishes the effect of TNF-α on nociceptive excitability. A, left: voltage dependence of slow inactivation of TTX-r sodium currents, expressed as peaks of TTX-r sodium currents during a step to 0 mV, normalized to Imax (fraction available, means ± SE) and plotted against conditioning voltages before (gray squares) and after application of TNF-α, together with lacosamide (open circles). Note: there was no change in voltage dependence of slow inactivation (see Table 2 for V1/2,slope values and statistical analysis, n = 8). A, right: the ratio of peak TTX-r sodium current before application of treatment (INa,control) to the peak current after the treatment (INa,treatment) with lacosamide and lacosamide + TNF-α. Values are means ± SE, n = 10; P = 0.32, compared with the control, one-sample t-test. B, left: voltage dependence of slow inactivation of TTX-r sodium currents, before (gray squares) and after application of TNF-α, together with brilliant blue G (BBG) (open circles). Note: there was no change in voltage dependence of slow inactivation (see Table 2 for V1/2,slope values and statistical analysis, n = 4). B, right: the ratio of peak TTX-r sodium current before application of treatment (INa,control) to the peak current after the treatment (INa,treatment) with BBG and BBG + TNF-α. Values are means ± SE, n = 4, P = 0.6, compared with the control, one-sample t-test. C, top: representative traces of current clamp recordings from the same nociceptor-like DRG neuron, following 500 pA, 1.5-s depolarizing ramps in control condition (left trace), 10 min after application of TNF-α (middle trace) and 5 min after application of lacosamide on top of 10 min application of TNF-α (right trace). See Table 4 for the numerical data and statistical analysis. Bottom: stimulation protocol. D: latency for paw withdrawal from hot plate (52°C), normalized to baseline, after intraplantar injection of saline (black squares); TNF-α alone (open circles); lacosamide alone (gray triangles); or TNF-α applied together with lacosamide (light gray inverted triangles). The values are means ± SE. Time point “0” refers to the baseline. Gray asterisks, comparison between the TNF-α group to TNF-α + Lacosamide group; ns, P > 0.05, *P < 0.05, two-way ANOVA followed by Bonferroni posttest, n = 6 for each group. E: mechanical threshold for paw withdrawal, normalized to baseline, obtained by von Frey filaments. The values are means ± SE, n = 6 for each group. Time point “0” refers to the baseline. Symbols are as in E; **P < 0.01.
Fig. 11.
Fig. 11.
TNF-α relieves TTX-r channels from ultraslow inactivation at hyperpolarized potentials. A fraction of channels available for activation (peaks of TTX-r sodium currents during a step to 0 mV, normalized to Imax, means ± SE) following prolonged 30-s holding at voltages, as indicated in the x-axis, before (gray squares) and 3 min after application of TNF-α (black circles). Currents were elicited by 10-ms test steps to 0 mV, after 30-s conditioning pulses (Vcond) to the indicated voltages. Note that the substantial current reduction, following prolonged holding at Vcond of −70 mV, −60 mV and −50 mV (black asterisks, comparison within Before group, ns, P > 0.05, **P < 0.01, ***P < 0.001 one-way ANOVA with post hoc Bonferroni, n = 6), was diminished by application of TNF-α (gray asterisks, comparison within TNF-α group, ns, P > 0.05, ***P < 0.001, one-way ANOVA with post hoc Bonferroni, n = 6). Note also that TNF-α leads to increase of available channels at −70 mV by about 15%. Light great asterisks, comparison between Before and TNF-α groups. *P < 0.05; **P < 0.01, two-way ANOVA with post hoc Bonferroni, n = 6. Inset: stimulation protocol.
Fig. 12.
Fig. 12.
Scheme depicting TNF-α-mediated nociceptive hyperexcitability. Left: in normal conditions, only a fraction of Nav 1.8 channels, which underlie slow TTX-r sodium current, are available for activation, due to slow inactivation (denoted by collapsed outer pore of the channel). This, together with the voltage dependence of activation of Nav 1.9 channels, which underlies persistent TTX-r sodium current, defines a threshold for AP initiation and the ability to fire repetitively after threshold is achieved. TNFR, TNF receptor. Right: TNF-α binds to its receptor and, via p38 MAPK, leads to a hyperpolarization shift in the activation of Nav 1.9 channels. Moreover, TNF-α-mediated activation of p38 MAPK, probably via direct phosphorylation of Nav 1.8 channels (see Hudmond et al. 2008 and also Fig. 9 here) increases the availability of Nav 1.8 channels, by modulating their slow inactivation. These processes lead to decrease in threshold, such that previously subthreshold stimuli now evoke APs, and to increase in gain, enhancing the ability of nociceptive neurons to fire repetitively, altogether producing nociceptive hyperexcitability.

References

    1. Abdulla FA, Smith PA. Axotomy- and autotomy-induced changes in the excitability of rat dorsal root ganglion neurons. J Neurophysiol 85: 630–643, 2001.
    1. Amaya F, Wang H, Costigan M, Allchorne AJ, Hatcher JP, Egerton J, Stean T, Morisset V, Grose D, Gunthorpe MJ, Chessell IP, Tate S, Green PJ, Woolf CJ. The voltage-gated sodium channel Na(v)1.9 is an effector of peripheral. J Neurosci 26: 12852–12860, 2006.
    1. Andres C, Hasenauer J, Ahn HS, Joseph EK, Isensee J, Theis FJ, Allgower F, Levine JD, Dib-Hajj SD, Waxman SG, Hucho T. Wound-healing growth factor, basic FGF, induces Erk1/2-dependent mechanical hyperalgesia. Pain 154: 2216–2226, 2013.
    1. Baker MD. Protein kinase C mediates up-regulation of tetrodotoxin-resistant, persistent Na+ current in rat and mouse sensory neurones. J Physiol 567: 851–867, 2005.
    1. Baker MD, Chandra SY, Ding Y, Waxman SG, Wood JN. GTP-induced tetrodotoxin-resistant Na+ current regulates excitability in mouse and rat small diameter sensory neurones. J Physiol 548: 373–382, 2003.
    1. Basbaum AI, Bautista DM, Scherrer G, Julius D. Cellular and molecular mechanisms of pain. Cell 139: 267–284, 2009.
    1. Bean BP. The action potential in mammalian central neurons. Nat Rev Neurosci 8: 18–20, 2007.
    1. Binshtok AM, Wang H, Zimmermann K, Amaya F, Vardeh D, Shi L, Brenner GJ, Ji RR, Bean BP, Woolf CJ, Samad TA. Nociceptors are interleukin-1beta sensors. J Neurosci 28: 14062–14073, 2008.
    1. Black JA, Waxman SG. Molecular identities of two tetrodotoxin-resistant sodium channels in corneal axons. Exp Eye Res 75: 193–199, 2002.
    1. Blair NT, Bean BP. Role of tetrodotoxin-resistant Na current slow inactivation in adaptation of action potential firing in small-diameter dorsal root ganglion neurons. J Neurosci 23: 10338–10350, 2003.
    1. Blair NT, Bean BP. Roles of tetrodotoxin (TTX)-sensitive Na current, TTX-resistant Na current, and Ca current in the action potentials of nociceptive sensory neurons. J Neurosci 22: 10277–10290, 2002.
    1. Bocksteins E, Raes AL, Van de Vijver G, Bruyns T, Van Bogaert PP, Snyders DJ. Kv2.1 and silent Kv subunits underlie the delayed rectifier K+ current in cultured small mouse DRG neurons. Am J Physiol Cell Physiol 296: C1271–C1278, 2009.
    1. Cardenas CG, Del Mar LP, Cooper BY, Scroggs RS. 5HT4 receptors couple positively to tetrodotoxin-insensitive sodium channels in a subpopulation of capsaicin-sensitive rat sensory neurons. J Neurosci 17: 7181–7189, 1997.
    1. Cardenas CG, Del Mar LP, Scroggs RS. Variation in serotonergic inhibition of calcium channel currents in four types of rat sensory neurons differentiated by membrane properties. J Neurophysiol 74: 1870–1879, 1995.
    1. Catterall WA, Goldin AL, Waxman SG. International Union of Pharmacology. XLVII nomenclature and structure-function. Pharmacol Rev 57: 397–409, 2005.
    1. Chance FS, Abbott LF, Reyes AD. Gain modulation from background synaptic input. Neuron 35: 773–782, 2002.
    1. Cummins TR, Black JA, Dib-Hajj SD, Waxman SG. Glial-derived neurotrophic factor upregulates expression of functional SNS and NaN sodium channels and their currents in axotomized dorsal root ganglion neurons. J Neurosci 20: 8754–8761, 2000.
    1. Cummins TR, Dib-Hajj SD, Black JA, Akopian AN, Wood JN, Waxman SG. A novel persistent tetrodotoxin-resistant sodium current in SNS-null and wild-type small primary sensory neurons. J Neurosci 19: RC43, 1999.
    1. 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 17: 3503–3514, 1997.
    1. Czeschik JC, Hagenacker T, Schafers M, Busselberg D. TNF-alpha differentially modulates ion channels of nociceptive neurons. Neurosci Lett 434: 293–298, 2008.
    1. Dib-Hajj SD, Cummins TR, Black JA, Waxman SG. Sodium channels in normal and pathological pain. Annu Rev Neurosci 33: 325–347, 2010.
    1. Du J, Haak LL, Phillips-Tansey E, Russell JT, McBain CJ. Frequency-dependent regulation of rat hippocampal somato-dendritic excitability by the K+ channel subunit Kv2.1. J Physiol 522: 19–31, 2000.
    1. Dubin AE, Patapoutian A. Nociceptors: the sensors of the pain pathway. J Clin Invest 120: 3760–3772, 2010.
    1. England S, Bevan S, Docherty RJ. PGE2 modulates the tetrodotoxin-resistant sodium current in neonatal rat dorsal root ganglion neurones via the cyclic AMP-protein kinase A cascade. J Physiol 495: 429–440, 1996.
    1. Flake NM, Lancaster E, Weinreich D, Gold MS. Absence of an association between axotomy-induced changes in sodium currents and excitability in DRG neurons from the adult rat. Pain 109: 471–480, 2004.
    1. Fleidervish IA, Gutnick MJ. Kinetics of slow inactivation of persistent sodium current in layer V neurons of mouse neocortical slices. J Neurophysiol 76: 2125–2130, 1996.
    1. Gold MS, Levine JD, Correa AM. Modulation of TTX-R INa by PKC and PKA and their role in PGE2-induced sensitization of rat sensory neurons in vitro. J Neurosci 18: 10345–10355, 1998.
    1. Gold MS, Reichling DB, Shuster MJ, Levine JD. Hyperalgesic agents increase a tetrodotoxin-resistant Na+ current in nociceptors. Proc Natl Acad Sci U S A 93: 1108–1112, 1996.
    1. Goldin A. Mechanisms of sodium channel inactivation. Curr Opin Neurobiol 13: 284–290, 2003.
    1. Hagenacker T, Czeschik JC, Schafers M, Busselberg D. Sensitization of voltage activated calcium channel currents for capsaicin in nociceptive neurons by tumor-necrosis-factor-alpha. Brain Res Bull 81: 157–163, 2010.
    1. Herzog RI, Cummins TR, Waxman SG. Persistent TTX-resistant Na+ current affects resting potential and response to depolarization in simulated spinal sensory neurons. J Neurophysiol 86: 1351–1364, 2001.
    1. Hines ML, Carnevale NT. The NEURON simulation environment. Neural Comput 9: 1179–1209, 1997.
    1. Huang J, Yang Y, Zhao P, Gerrits MM, Hoeijmakers JG, Bekelaar K, Merkies IS, Faber CG, Dib-Hajj SD, Waxman SG. Small-fiber neuropathy Nav1.8 mutation shifts activation to hyperpolarized potentials and increases excitability of dorsal root ganglion neurons. J Neurosci 33: 14087–14097, 2013.
    1. Huang ZJ, Song XJ. Differing alterations of sodium currents in small dorsal root ganglion neurons after ganglion compression and peripheral nerve injury. Mol Pain 4: 20–35, 2008.
    1. Hucho T, Levine JD. Signaling pathways in sensitization: toward a nociceptor cell biology. Neuron 55: 365–376, 2007.
    1. Hudmon A, Choi JS, Tyrrell L, Black JA, Rush AM, Waxman SG, Dib-Hajj SD. Phosphorylation of sodium channel Na(v)18 by p38 mitogen-activated protein kinase increases current density in dorsal root ganglion neurons. J Neurosci 28: 3190–3201, 2008.
    1. Jenerick H. Phase plane trajectories of the muscle spike potential. Biophys J 3: 363–377, 1963.
    1. Jin X, Gereau RW. Acute p38-mediated modulation of tetrodotoxin-resistant sodium channels in mouse sensory neurons by tumor necrosis factor-alpha. J Neurosci 26: 246–255, 2006.
    1. Jo S, Bean BP. Inhibition of neuronal voltage-gated sodium channels by brilliant blue G. Mol Pharmacol 80: 247–257, 2011.
    1. Junger H, Sorkin LS. Nociceptive and inflammatory effects of subcutaneous TNFalpha. Pain 85: 145–151, 2000.
    1. Kispersky TJ, Caplan JS, Marder E. Increase in sodium conductance decreases firing rate and gain in model neurons. J Neurosci 32: 10995–11004, 2012.
    1. Leffler A, Cummins TR, Dib-Hajj SD, Hormuzdiar WN, Black JA, Waxman SG. GDNF and NGF reverse changes in repriming of TTX-sensitive Na(+) currents following axotomy of dorsal root ganglion neurons. J Neurophysiol 88: 650–658, 2002.
    1. Li Y, Ji A, Weihe E, Schafer MK. Cell-specific expression and lipopolysaccharide-induced regulation of tumor necrosis factor alpha (TNFalpha) and TNF receptors in rat dorsal root ganglion. J Neurosci 24: 9623–9631, 2004.
    1. Liu BG, Dobretsov M, Stimers JR, Zhang JM. Tumor necrosis factor-alpha suppresses activation of sustained potassium currents in rat small diameter sensory neurons. Open Pain J 1: 1–16, 2008.
    1. Maier JA, Hla T, Maciag T. Cyclooxygenase is an immediate-early gene induced by interleukin-1 in human endothelial cells. J Biol Chem 265: 10805–10808, 1990.
    1. Maingret F, Coste B, Padilla F, Clerc N, Crest M, Korogod SM, Delmas P. Inflammatory mediators increase Nav1.9 current and excitability in nociceptors through a coincident detection mechanism. J Gen Physiol 131: 211–225, 2008.
    1. Mannion RJ, Costigan M, Decosterd I, Amaya F, Ma QP, Holstege JC, Ji RR, Acheson A, Lindsay RM, Wilkinson Ga, Woolf CJ. Neurotrophins: peripherally and centrally acting modulators of tactile stimulus-induced inflammatory pain hypersensitivity. Proc Natl Acad Sci U S A 96: 9385–9390, 1999.
    1. 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 154: 69–82, 1978.
    1. McCarthy PW, Lawson SN. Differing action potential shapes in rat dorsal root ganglion neurones related to their substance P and calcitonin gene-related peptide immunoreactivity. J Comp Neurol 388: 541–549, 1997.
    1. Nicol GD, Lopshire JC, Pafford CM. Tumor necrosis factor enhances the capsaicin sensitivity of rat sensory neurons. J Neurosci 17: 975–982, 1997.
    1. O'Connell KM, Loftus R, Tamkun MM. Localization-dependent activity of the Kv2.1 delayed-rectifier K+ channel. Proc Natl Acad Sci U S A 107: 12351–12356, 2010.
    1. Ogata N, Tatebayashi H. Kinetic analysis of two types of Na+ channels in rat dorsal root ganglia. J Physiol 466: 9–37, 1993.
    1. Ogata N, Tatebayashi H. Slow inactivation of tetrodotoxin-insensitive Na+ channels in neurons of rat dorsal root ganglia. J Membr Biol 129: 71–80, 1992.
    1. Ogawa T, Hayashi T, Kyoizumi S, Kusunoki Y, Nakachi K, MacPhee DG, Trosko JE, Kataoka K, Yorioka N. Anisomycin downregulates gap-junctional intercellular communication via the p38 MAP-kinase pathway. J Cell Sci 117: 2087–2096, 2004.
    1. Ostenfeld T, Krishen A, Lai RY, Bullman J, Baines AJ, Green J, Anand P, Kelly M. Analgesic efficacy and safety of the novel p38 MAP kinase inhibitor, losmapimod, in patients with neuropathic pain following peripheral nerve injury: a double-blind, placebo-controlled study. Eur J Pain 17: 844–857, 2013.
    1. Persson AK, Black JA, Gasser A, Cheng X, Fischer TZ, Waxman SG. Sodium-calcium exchanger and multiple sodium channel isoforms in intra-epidermal nerve terminals. Mol Pain 6: 84, 2010.
    1. Pollock J, McFarlane SM, Connell MC, Zehavi U, Vandenabeele P, MacEwan DJ, Scott RH. TNF-alpha receptors simultaneously activate Ca2+ mobilisation and stress kinases. Neuropharmacology 42: 93–106, 2002.
    1. Prescott SA, De Koninck Y. Gain control of firing rate by shunting inhibition: roles of synaptic noise and dendritic saturation. Proc Natl Acad Sci U S A 100: 2076–2081, 2003.
    1. Prescott SA, Sejnowski TJ, De 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 2: 32, 2006.
    1. Ramachandra R, McGrew SY, Baxter JC, Kiveric E, Elmslie KS. Tetrodotoxin-resistant voltage-dependent sodium channels in identified muscle afferent neurons. J Neurophysiol 108: 2230–2241, 2012.
    1. Ritter DM, Ho C, O'Leary ME, Covarrubias M. Modulation of Kv3.4 channel N-type inactivation by protein kinase C shapes the action potential in dorsal root ganglion neurons. J Physiol 590: 145–161, 2012.
    1. Roy ML, Narahashi T. Differential properties of tetrodotoxin-sensitive and tetrodotoxin-resistant sodium channels in rat dorsal root ganglion neurons. J Neurosci 12: 2104–2111, 1992.
    1. Rush AM, Cummins TR, Waxman SG. Multiple sodium channels and their roles in electrogenesis within dorsal root ganglion neurons. J Physiol 579: 1–14, 2007.
    1. Safieh-Garabedian B, Kanaan SA, Haddad JJ, Jaoude PA, Jabbur SJ, Saade NE. Involvement of interleukin-1 beta, nerve growth factor and prostaglandin E2 in endotoxin-induced localized inflammatory hyperalgesia. Br J Pharmacol 121: 1619–1626, 1997.
    1. Scroggs RS. Serotonin upregulates low- and high-threshold tetrodotoxin-resistant sodium channels in the same subpopulation of rat nociceptors. Neuroscience 165: 1293–1300, 2010.
    1. Scroggs RS. Up-regulation of low-threshold tetrodotoxin-resistant Na+ current via activation of a cyclic AMP/protein kinase A pathway in nociceptor-like rat dorsal root ganglion cells. Neuroscience 186: 13–20, 2011.
    1. Sheets PL, Heers C, Stoehr T, Cummins TR. Differential block of sensory neuronal voltage-gated sodium channels by lacosamide [(2R)-2-(acetylamino)-N-benzyl-3-methoxypropanamide], lidocaine, and carbamazepine. J Pharmacol Exp Ther 326: 89–99, 2008.
    1. Smith MT, Woodruff TM, Wyse BD, Muralidharan A, Walther T. A small molecule angiotensin II type 2 receptor [AT(2)R] antagonist produces analgesia in a rat model of neuropathic pain by inhibition of p38 mitogen-activated protein kinase (MAPK) and p44/p42 MAPK activation in the dorsal root ganglia. Pain Med 14: 1557–1568, 2013.
    1. Sorkin LS, Doom CM. Epineurial application of TNF elicits an acute mechanical hyperalgesia in the awake rat. J Peripher Nerv Syst 5: 96–100, 2000.
    1. Sorkin LS, Xiao WH, Wagner R, Myers RR. Tumour necrosis factor-alpha induces ectopic activity in nociceptive primary afferent fibres. Neuroscience 81: 255–262, 1997.
    1. Tal M, Bennett GJ. Extra-territorial pain in rats with a peripheral mononeuropathy. Pain 57: 375–382, 1994.
    1. Todt H, Dudley SC Jr, Kyle JW, French RJ, and Fozzard HA. Ultra-slow inactivation in mu1 Na+ channels is produced by a structural rearrangement of the outer vestibule. Biophys J 76: 1335–1345, 1999.
    1. Tripathi PK, Cardenas CG, Cardenas CA, Scroggs RS. Up-regulation of tetrodotoxin-sensitive sodium currents by prostaglandin E(2) in type-4 rat dorsal root ganglion cells. Neuroscience 185: 14–26, 2011.
    1. Tripathi PK, Trujillo L, Cardenas Ca Cardenas CG, de Armendi AJ, Scroggs RS. Analysis of the variation in use-dependent inactivation of high-threshold tetrodotoxin-resistant sodium currents recorded from rat sensory neurons. Neuroscience 143: 923–938, 2006.
    1. Tsantoulas C, Zhu L, Yip P, Grist J, Michael GJ, McMahon SB. Kv2 dysfunction after peripheral axotomy enhances sensory neuron responsiveness to sustained input. Exp Neurol 251: 115–126, 2014.
    1. Wittmack EK, Rush AM, Hudmon A, Waxman SG, Dib-Hajj SD. Voltage-gated sodium channel Nav16 is modulated by p38 mitogen-activated protein kinase. J Neurosci 25: 6621–6630, 2005.
    1. Woolf CJ, Allchorne A, Safieh-Garabedian B, Poole S. Cytokines, nerve growth factor and inflammatory hyperalgesia: the contribution of tumour necrosis factor alpha. Br J Pharmacol 121: 417–424, 1997.
    1. Woolf CJ, Ma Q. Nociceptors–noxious stimulus detectors. Neuron 55: 353–364, 2007.
    1. Zarrabi T, Cervenka R, Sandtner W, Lukacs P, Koenig X, Hilber K, Mille M, Lipkind GM, Fozzard HA, Todt H. A molecular switch between the outer and the inner vestibules of the voltage-gated Na+ channel. J Biol Chem 285: 39458–39470, 2010.
    1. Zimmermann K, Leffler A, Babes A, Cendan CM, Carr RW, Kobayashi Ji Nau C, Wood JN, Reeh PW. Sensory neuron sodium channel Nav1.8 is essential for pain at low temperatures. Nature 447: 855–858, 2007.

Source: PubMed

3
S'abonner