Dynamic-Clamp Analysis of Wild-Type Human NaV1.7 and Erythromelalgia Mutant Channel L858H*

Dmytro V. Vasylyev, Chongyang Han, Peng Zhao, Sulayman Dib-Hajj, and Stephen G. Waxman

The link between sodium channel NaV1.7 and pain has been strengthened by identification of gain-of-function mutations in patients with inherited erythromelalgia (IEM), a genetic model of neuropathic pain in humans. A firm mechanistic link to nociceptor dysfunction has been precluded because assessments of the effect of the mutations on nociceptor function have thus far depended on electrophysiological recordings from dorsal root ganglia (DRG) neurons transfected with wild-type (WT) or mutant NaV1.7 channels, which do not permit accurate calibration of the level of NaV1.7 channel expression. Here, we report an analysis of the function of WT NaV1.7 and IEM L858H mutation within small DRG neurons using dynamic-clamp. We describe the functional relationship between current threshold for action potential generation and the level of WT NaV1.7 conductance in primary nociceptive neurons and demonstrate the basis for hyperexcitability at physiologically relevant levels of L858H channel conductance. We demonstrate that the L858H mutation, when modeled using dynamic-clamp at physiological levels within DRG neurons, produces a dramatically enhanced persistent current, resulting in 27-fold amplification of net sodium influx during subthreshold depolarizations and even greater amplification during interspike intervals, which provide a mechanistic basis for reduced current threshold and enhanced action potential firing probability. These results show, for the first time, a linear correlation between the level of NaV1.7 conductance and current threshold in DRG neurons. Our observations demonstrate changes in sodium influx that provide a mechanistic link between the altered biophysical properties of a mutant NaV1.7 channel and nociceptor hyperexcitability underlying the pain phenotype in IEM.

The NaV voltage-gated sodium channel is pre­ferentially expressed in primary nociceptors (Persson et al. 2010; Rush et al. 2007; Toledo-Aral et al. 1997) and activates at subthreshold membrane voltages so as to boost membrane responses to small depolarizing stimuli both in neuronal somata (Ahn et al. 2013; Cummins et al. 1998; Herzog et al. 2003; Rush et al. 2007) and at axon endings of small dorsal root ganglia (DRG) neurons (Vasylyev and Waxman 2012). Inherited erythromelalgia (IEM), the first human pain disorder linked to a sodium channel, is widely regarded as a genetic model of neuropathic pain. When studied in a mammalian heterologous expression system, the L858H mutation, one of the first NaV1.7 mutations linked to IEM (Cummins et al. 2004; Rush et al. 2006; Yang et al. 2004), produces a hyperpolarizing shift in channel activation and increases the amplitude of the response of the channel to slow, small depolarizations. When small DRG neurons transfected with L858H and wild-type (WT) NaV1.7 channels were studied by current-clamp, L858H was found to render DRG neurons hyperexcitable in a manner consistent with the severe pain characteristic of the erythromelalgia phenotype (Rush et al. 2006).

However, a mechanistic link between altered biophysical properties of mutant NaV1.7 channels and hyperexcitability of primary nociceptor neurons carrying these mutant channels has not yet been established. In the present study, we used dynamic-clamp (Kemenes et al. 2011; Samu et al. 2012; Sharp et al. 1993) recordings to titrate levels of NaV1.7 WT and L858H conductances so that we could study the effect of precisely calibrated, physiologically relevant levels of conductance on small DRG neuron excitability. This approach permitted us to vary the ratio of expression levels of WT and L858H within the physiological range within single neurons. This is important because, although the exact ratio of WT and L858H channels is not known, it is reasonable to suggest a 1:1 ratio of functional expressions of WT and L858H mutant allele in the affected individual (Yang et al. 2004). Using dynamic-clamp, we show that action potential (AP) current threshold of DRG neurons is regulated by NaV1.7 conductance in a remarkably linear manner. We also show, in small DRG neurons in which we inserted the L858H mutant by dynamic-clamp, that the mutation substantially increases channel activity at membrane potentials below the AP threshold, producing a large persistent current that depolarizes the membrane potential. We also show that sodium influx via L858H mutant channels is increased not only during subthreshold membrane depolarizations, but also during interspike intervals. These observations on small DRG neurons within a physiologically relevant range of levels of conductance provide a quantitative mechanistic basis for understanding the role of WT NaV1.7 in healthy DRG neurons and the enhanced excitability of primary nociceptors expressing L858H channels that underlies the pain phenotype in humans carrying this mutation.

Materials and Methods

Voltage-Clamp Recordings of NaV1.7

Human NaV1.7 channels were stably expressed in HEK-293 cell line (Cummins et al. 1998). Pipettes were pulled from glass capillaries (cat. no. PG10165-4; World Precision Instruments, Sarasota, FL) and had resistance 1.5–2 MD when filled with the intracellular solution in mM: 140 CsCl, 10 NaCl, 0.5 EGTA, 10 HEPES, 3 MgATP, 10 glucose, pH 7.3 with CsOH. The extracellular solution was HBSS (cat. no. 14025; Invitrogen) in mM: 1.3 CaCl2, 0.5 MgCl2, 0.4 MgSO4, 5.3 KCl, 0.4 KH2PO4, 4.2 NaHCO3, 138 NaCl, 0.3 Na2HPO4, 5.6 glucose, supplemented with 15 mM NaCl (320 mosM). The liquid junction potential (+3.7 mV) between pipette and bath solutions was measured according to Neher (1992) and was not compensated. Whole cell voltage-clamp recordings were made using Axopatch 200B amplifier. Currents were low-pass filtered at 10 kHz and digitized/stored at 100 kHz by Digidata 1440A DAC using pCLAMP 10 software (Molecular Devices, Sunnyvale, CA). The series resistance was compensated by 80–85%. -P/4 protocol was used for current-voltage (IV) and deactivation protocols to subtract uncompensated leak and capacitive currents. Recordings were made at room temperature (23–24°C). All data are presented as means ± SE. Data were analyzed using pCLAMP 10 and Origin 8.5 (OriginLab, Northampton, MA) software.

Assessment of NaV1.7 Contribution to TTX-S Current in Small DRG Neurons

For assessment of the contribution of NaV1.7 to the TTX-sensitive (TTX-S) current, DRG neurons were isolated from WT or NaV1.7-knockout (KO) mice (11–16 wk old) and cultured as previously described (Dib-Hajj et al. 2009). Whole cell voltage-clamp recordings of small (20–25 µm) DRG neurons were obtained at room temperature (20–22°C) within 2–8 h in culture using an EPC 10 amplifier (HEKA Electronics) and fire-polished electrodes (1–2 MD) fabricated from 1.6-mm-outer-diameter borosilicate glass micropipettes (World Precision Instruments). The pipette potential was adjusted to zero before seal formation, and liquid junction potential was not corrected. Voltage errors were minimized with 80–90% series resistance compensation, and linear leak currents and capacitance artifacts were subtracted out using the P/6 method. Currents were acquired with PULSE software (HEKA Electronics) 5 min after establishing whole cell configuration, sampled at a rate of 50 kHz, and filtered at 2.9 kHz. The pipette solution contained in mM: 140 CsF, 10 NaCl, 1 EGTA, and 10 HEPES, pH 7.3 with CsOH (adjusted to 315 mosM with dextrose). The extracellular bath solution contained in mM: 70 NaCl, 70 choline chloride, 3 KCl, 1 MgCl2, 1 CaCl2, 20 TEACl, 5 CsCl, 1 4-AP, 0.1 CdCl2, and 10 HEPES, pH 7.31 with NaOH (326 mosM). The amplitude of TTX-S sodium current was estimated by two protocols as previously described (Rush et al. 2005). Cells were first CsCl, 1 4-AP, 0.1 CdCl held at –80 mV, and a 500-ms depolarizing prepulse to –50 mV was applied to inactivate the TTX-S sodium channels while leaving NaV1.8 current intact, followed by a series of step depolarizations from –70 to +20 mV (in 5-mV increments). Cells exhibiting NaV1.9 currents were excluded from data analysis. For the second protocol, cells were held at –80 mV, a 500-ms hyperpolarizing prepulse to –120 mV was applied to rescue the TTX-S sodium channels from inactivation, and total sodium currents were evoked by a series of depolarizing steps from –70 to +20 mV (in 5-mV increments). The TTX-S sodium current was obtained by subtraction of the currents obtained from the two protocols.

Dynamic-Clamp Recording

DRG neurons (soma diameter 21–26 µm; 24.4 ± 0.3, n = 25) obtained from postnatal day 05 Sprague-Dawley rats were grown in primary culture for 2–3 days (Ahn et al. 2013; Dib-Hajj et al. 2009). Small DRG neurons were dynamically clamped (Kemenes et al. 2011; Samu et al. 2012; Sharp et al. 1993) in whole cell configuration using patch pipettes pulled from glass capillaries (cat. no. PG10165-4; World Precision Instruments). Pipette resistance was 1.5–2 MD when filled with the intracellular solution in mM: 140 KCl, 3 MgATP, 0.5 EGTA, 5 HEPES, 10 glucose, pH 7.3 with KOH (adjusted to 325 mosM with sucrose). The extracellular solution was HBSS (cat. no. 14025; Invitrogen) in mM: 1.3 CaCl2, 0.5 MgCl2, 0.4 MgSO4, 5.3 KCl, 0.4 KH2PO4, 4.2 NaHCO3, 138 NaCl, 0.3 Na2HPO4, 5.6 glucose (adjusted to 325 mosM with sucrose). Liquid junction potential (+3.8 mV) between pipette and bath solutions was not compensated. Membrane voltages and currents were recorded in dynamic-clamp using MultiClamp 700B amplifier (Molecular Devices) interfaced with CED Power1401 mk II DAI and Signal software (Cambridge Electronic Design, Cambridge, United Kingdom), digitized by Digidata 1440A DAC, and stored on the hard disk using pCLAMP 10 software. Capacitance neutralization and bridge balance were rigorously employed to minimize the effect of electrode capacitance and series resistance on the dynamic-clamp recordings. IV traces were filtered at 10 kHz and digitized at 50 kHz. Series resistance was compensated by 80–85%. Endogenous sodium current recorded in whole cell voltage-clamp was evoked from a holding potential of –50 mV by a test pulse to –10 mV with or without preceding 0.5-s prepulse to –100 mV to remove steady-state inactivation of TTX-S channels. TTX-S channels are mainly inactivated at –50 mV, whereas NaV1.8 channels are still available for activation at this potential because steady-state inactivation of NaV1.8 is shifted in a depolarized direction (Catterall et al. 2005). Thus, as described by Cummins and Waxman (1997) and Rush et al. (2005), as a measure of total TTX-S current, we used a conditioning/subtraction protocol and assessed the peak current on the trace obtained by digital subtraction of the current trace evoked by the test voltage without preceding prepulse from that with the prepulse, respectively.

We estimated the contribution of NaV1.7 to the total TTX-S current in our assay based on measurements of the total TTX-S current amplitude in WT and NaV1.7-KO DRG neurons (see above). To evaluate the effect of defined levels of L858H functional expression on excitability of small DRG neuron using dynamic-clamp recording, we first estimated the endogenous NaV1.7 conductance and then substituted graded amounts of endogenous conductance by equal amounts of L858H channel conductance, thus achieving a specified substitution ratio (SR) via a dynamic-clamp operation, SR × gmax(L858H – WT), where gmax is maximal NaV1.7 endogenous conductance.

Recordings were made at room temperature (23–24°C). All data are presented as means ± SE. Data were analyzed using pCLAMP 10 and Origin 8.5 software. Unless specified otherwise, the hypothesis that population means are significantly different was checked using Mann-Whitney nonparametric test (P < 0.05, P < 0.01, and P < 0.001).

Kinetic Model of NaV1.7 Channel

We developed our model of NaV1.7 channel using Hodgkin–Huxley equations dm/dt = αm(1 − m) − ßmm; dh/dt = αh(1 − h) − ßhh, where m and h are channel activation and inactivation variable and α (ß) are forward (backward) rate constants, respectively. Channel states were independent with a first-order reaction between states. Thus channel activation and deactivation were considered as transitions between closed and open states, whereas channel inactivation and repriming were assumed to be transitions between primed and inactivated states, respectively. NaV1.7 channel steady-state parameters and kinetics obtained based on electrophysiological recordings were converted into appropriate rate constants at respective voltages using the equations α = m/τ, ß = (1 − m)/τ. These reaction rate constants were fitted with Boltzmann equations of the form y = A2 + (A1 − A2)/{1 + exp[(VV1/2)/k]}, where V is membrane voltage, V1/2 is voltage when reaction rate is half-maximal, and k is slope coefficient. Fits were converted into steady-state inactivation (activation) variables and inactivation (activation) time constants according to m = α/( α + ß) and τ = 1/(α + ß). The latter curves were overplayed on the experimental data to provide feedback to the rate constants fitting step. This cycle was manually repeated until the best possible fit of the experimental data was achieved. We obtained the following rate constants for the WT NaV1.7 channel model:

αm = 10.22 − 10.22/{1 + exp[(V + 7.19)/15.43]}, ßm = 23.76/{1 + exp[(V + 70.37)/14.53]};

αh = 0.0744/{1 + exp[(V + 99.76)/11.07]}, ßh = 2.54 − 2.54/{1 + exp[(V + 7.8)/10.68]}.

We modeled the L858H IEM mutation because it has been well-studied at the voltage-clamp (Cummins et al. 2004) and current-clamp (Rush et al. 2006) levels. We focused on altered activation of the mutant channels because a hyperpolarizing shift in activation is common to all IEM mutant channels (Dib-Hajj et al. 2010). We did not model slow inactivation in the mutant channel because we limited stimulation to 1-s trains at 10 Hz where the development of slow inactivation is not appreciable. The L858H NaV1.7 channel model was described by the following equations:

αm = 9.1 − 9.1/{1 + exp[(V + 11.52)/22.49]}, ßm = 23.76/{1 + exp[(V + 87.6)/14.53]};

αh = 0.0744/{1 + exp[(V + 99.76)/11.07]}, ßh = 2.54 − 2.54/{1 + exp[(V + 7.8)/10.68]}.

Sodium current was described by INa = gmax × m × h × (VmENa), where Vm is membrane voltage potential and ENa = 65 mV is sodium reversal potential. Currents evoked by different voltage protocols were calculated in 10-µs precision using a custom program written in Origin 8.5 LabTalk.

NaV1.7 current kinetics in response to square test pulses and the resulting IV curve were identical either when calculated in LabTalk or obtained from dynamic-clamp recordings on a vendors-supplied physical cell model. All data are presented as means ± SE. Data were analyzed using pCLAMP 10 and Origin 8.5 software.

Results

Kinetic Model of WT NaV1.7 Based on Hodgkin–Huxley Equations

Channel kinetics were described based on a Hodgkin–Huxley model of sodium channel with several independent states and a first-order reaction between the states. Channel activation and deactivation were considered as transitions between closed and open states, whereas channel inactivation and repriming were assumed to be transitions between primed (open or closed) and inactivated states (Hille 1978; Hodgkin and Huxley 1952).

NaV1.7 currents were recorded in whole cell voltage-clamp to obtain channel steady-state and kinetics parameters. Steady-state inactivation (h) was calculated as the ratio of peak current amplitude to the maximal peak current amplitude elicited by a 0-mV test voltage following a 1-s prepulse at different voltages ranging from –100 to –40 mV in 5-mV increments from a holding potential of –100 mV (figure 1A). Steady-state activation (m) was defined as the ratio of NaV1.7 g = Ipeak/(VmENa) determined at the respective membrane potentials to the NaV1.7 gmax (figure 1A). Steady-state relationships of recombinant NaV1.7 (n = 8 cells) were best fit using a Boltzmann equation with the following parameters: V1/2 = –73.9 mV, k = 6.2 mV and V1/2 = –20.4 mV, k = 7.2 mV for channel steady-state inactivation (n = 8) and steady-state activation (n = 17), respectively. Our model accurately described NaV1.7 channel steady-state properties as can be seen from the respective Boltzmann fits of the modeled steady-state curves, V1/2 = –72.6 mV, k = 5.7 mV and V1/2 = –20.4 mV, k = 7.1 mV (figure 1A).

10310_006b_fig_001.jpg

Figure 1 Kinetic model of NaV1.7 voltage-gated sodium channel based on Hodgkin–Huxley equations. (A) Voltage dependence of conductance/maximal conductance (G/Gmax) at steady-state for inactivation (squares; n = 8) and activation (circles; n = 17) of NaV1.7 channel. Solid lines are derived from the following equations: h = αh/(αh + ßh); (m)3 = [αm/(αm + ßm)]3, where m and h are channel activation and inactivation variables and α and ß are forward and backward rate constants (see Materials and Methods). (B) Voltage dependencies of activation (solid circles; n = 8) and deactivation (open circles; n = 14) time constants. Deactivation of the NaV1.7 current was fitted with a single exponential, whereas channel activation was fitted with a single exponential raised to the 3rd power. (C) Inactivation (solid squares; n = 13) and removal of inactivation (open squares; n = 7) time constant obtained from a single-exponential fit of the respective data. Inset shows data replotted with time constants on a log scale. (D) Time sequences of m3 and h variables along with the resulting open probability (m3h) obtained in the model response to a series of voltage steps ranged from –60 to 40 mV in 5-mV increments from a holding potential of –110 mV. (E) Rising phase of the NaV1.7 current (bottom) in response to a –30-mV test voltage (top) was fitted with a single exponential function of different powers. The best fit was obtained using 3rd-power exponential as determined based on the residual (middle); the residual of the 3rd-power exponential fit was not substantially different from the background noise, thus 4th-power exponential had not further improved the fit. (F and G) NaV1.7 current evoked by test pulses ranged from –50 to –25 mV in 5-mV increments (black) and overlaid on the m3h model (blue) at the respective voltages. Vertical dashed line denotes stimulus onset time. (H) Current traces (top) and the current-voltage (IV) curve (bottom) of NaV1.7 current (black) and the m3h model (blue) at the respective test voltages.

NaV1.7 current activated in a sigmoidal manner with an apparent delay (figure 1, E and F), which suggests that the channel undergoes multiple closed-state transitions before activation. Channel activation kinetics were best described by a single exponential raised to the third power (figure 1E). Channel activation (figure 1B; n = 8) was determined from current traces elicited by test voltages ranging from –55 to 60 mV in 5- to 10-mV increments applied from –100-mV holding potential. Deactivation time constant (figure 1B; n = 14) was determined from a single-exponential fit of the respective portions of current traces (“tail currents”) measured at different test voltages ranging from –100 to –20 mV following a 0.5-ms voltage step to –10 mV from a holding potential of –100 mV. The falling phase of the current (90 to 10% amplitude) was fitted with a single exponential to obtain the inactivation time constant (figure 1C; n = 7). The following protocol was used to evaluate channel repriming kinetics [for details, see Cummins et al. (1998)]. First, a 20-ms test voltage step to 0 mV was applied from a holding potential of –100 mV to inactivate the channel, and then repriming voltage steps of different durations (from 2 to 1,000 ms) were applied at a given voltage followed by the second 20-ms test voltage to 0 mV. The ratios of peak current amplitudes measured at the first and second test voltages were plotted as a function of the repriming step duration; this function was fitted with a single exponential to obtain a time constant of channel removal from inactivation (figure 1C; n = 14).

Our model effectively described channel kinetics as can be seen from the close match between the modeled and experimentally determined activation-deactivation (figure 1B) and inactivation-repriming (figure 1C) time constants at a wide range of physiological membrane voltages. The obtained Hodgkin–Huxley variables m3 and h were both highly voltage-dependent with submillisecond activation kinetics resulting in a transient channel open probability in response to a series of voltage steps (figure 1D). Our model accurately followed the kinetics of NaV1.7 current evoked by a series of depolarization steps (figure 1, F and G). The resulting IV relationships also provided a match between modeled and experimentally determined data at a wide range of physiological membrane voltages (figure 1H). Thereafter, we utilized this model (WT) and its modification for the L858H mutant channel to study mechanisms of functional contribution of NaV1.7 channel to neuronal excitability in normal and pathological conditions.

NaV1.7 Conductance Regulates Current Threshold for AP Generation in a Linear Manner

We performed a quantitative analysis of the effect of graded additions or subtractions of sodium conductance resulting from NaV1.7 channel function on small DRG neuron excitability using dynamic-clamp. Although we had the capability to transfect DRG neurons with NaV1.7 channels to study the effect of the channel on neuronal excitability (Dib-Hajj et al. 2009; Rush et al. 2006), transfection does not permit accurate calibration of the level of NaV1.7 channel that is being expressed. Thus we carried out dynamic-clamp recording in DRG neurons based on experimentally determined NaV1.7 gating properties and an assumption that NaV1.7 contributes on average 70% of the TTX-S current in small DRG neurons. We based our estimate of the NaV1.7 contribution of 70% of the TTX-S current in small DRG neurons on our measurements of total TTX-S currents in mice that are deficient in NaV1.7 (figure 2). Figure 2A shows representative family traces of TTX-S sodium currents recorded in DRG neurons from WT and NaV1.7-KO mice, respectively. Peak current densities were 143 ± 17 pA/pF (n = 16) for TTX-S sodium channels recorded from WT DRG neurons. However, DRG neurons from NaV1.7-KO mice produced significantly smaller TTX-S sodium currents with current densities ~32% (46 ± 9 pA/pF, n = 16) of that in WT DRG neurons (figure 2B).

10310_006b_fig_002.jpg

Figure 2 Measurement of NaV1.7 contribution to TTX-sensitive (TTX-S) current in small dorsal root ganglia (DRG) neurons. (A) Representative IV curve family traces of TTX-S sodium currents recorded from wild-type (WT) and NaV1.7-knockout (KO) DRG neurons, respectively. (B) TTX-S currents in NaV1.7-KO DRG neurons are significantly smaller than in WT DRG neurons. ***P < 0.001.

Using dynamic-clamp with the NaV1.7 channel model, we studied how increases in the level of NaV1.7 conductance affect small DRG neuron excitability. Dynamic-clamp allowed us to add graded increments of NaV1.7 current to the cell, matching the injected conductances to the predicted contribution of NaV1.7 channels to the TTX-S current (70% of total TTX-S current in each cell is assumed to be produced by NaV1.7; see above). We found that the current threshold (defined when the 2nd differential of AP changes its sign) inversely correlated in a linear fashion (r2 = 0.97, current threshold; r2 = 0.98, incremental threshold change) with the addition of NaV1.7 conductance (figure 3, A and B). Current threshold decreased from its original value (426 ± 82 pA, n = 9) to a value of 310 ± 65 pA (n = 9) when NaV1.7 was doubled by dynamic-clamp. Since current threshold variability was large (current threshold in 9 control cells was 350, 230, 470, 525, 340, 240, 90, 840, and 750 pA) in small DRG neurons, we normalized the effect of the NaV1.7 channel addition on current threshold and expressed it in the form 100% × ΔCT/CT0, where ΔCT is current threshold change at the respective level of NaV1.7 addition and CT0 is native current threshold. The normalized current threshold change significantly decreased (Mann-Whitney, P < 0.001) from –5.6 ± 0.8% (addition of 12.5% NaV1.7) to –28.1 ± 3.5% (addition of 100% NaV1.7). Importantly, stimulation with a current equivalent to the current threshold when NaV1.7 conductance was doubled did not elicit an AP in a DRG neuron with native levels of NaV1.7 (figure 3, A and B). Electronic addition of NaV1.7 conductance also resulted in an enhancement of AP firing probability (figure 3C). A 10-Hz train of 10 current pulses 10-ms duration each applied at the threshold level evoked 1.7 ± 0.3 APs in control cells vs. 4.3 ± 0.5 (n = 8; P < 0.01), 6.1 ± 0.7 (n = 8; P < 0.01), 8.4 ± 0.7 (n = 8; P < 0.001), and 9.4 ± 0.5 (n = 8; P < 0.001) after electronic addition of 12.5, 25, 50, and 100% NaV1.7 conductance, respectively (figure 3D).

10310_006b_fig_003.jpg

Figure 3 Additional NaV1.7 conductance lowers current threshold for action potential (AP) generation and increases AP firing probability. (A) 10-ms-long test pulses (top) applied at the original threshold without additional conductance (left) and at the threshold after electronic addition of 50% NaV1.7 conductance (right) elicited APs (bottom). Control stimulus and AP traces are shown in black, and those after addition of 50% NaV1.7 conductance are shown in blue. (B) Averages of current threshold (top) and current threshold change (bottom) plotted as a function of additional NaV1.7 conductance (n = 9). The solid line is a linear regression fit (r2 = 0.97, top; r2 = 0.97, bottom) of the data. Statistical analysis on the bottom was performed between threshold increments obtained at 12.5% and at the respective percentage of conductance increment. *P < 0.05, **P < 0.01, and ***P < 0.001. (C) APs evoked by a 1-s-long, 10-Hz train of current pulses (10-ms pulse width) applied at the original threshold level (100% native NaV1.7 conductance). Electronic addition of NaV1.7 conductance (expressed as the incremental increase over the endogenous NaV1.7 conductance) is noted on the y-axis. Scale bar, 200 ms. (D) Averages of the number of APs evoked by the protocol described in C plotted as a function of dynamically introduced NaV1.7 conductance (n = 8).

It is reasonable to suggest that electronic subtraction of NaV1.7 conductance will reduce neuronal excitability. Indeed, incremental addition of graded levels of negative NaV1.7 conductance resulted in an increase of current threshold in a linear fashion (r2 = 0.99) and in incremental (r2 = 0.98) reduction of AP firing probability (figure 4, A and B). A 10-Hz train of 10 current pulses applied at 1.5× threshold level evoked 9.9 ± 0.1 APs (n = 15) in control cells vs. 9.4 ± 0.4 (n = 15), 7.8 ± 0.6 (n = 15; P < 0.001), 5.3 ± 1.7 (n = 15; P < 0.001), and 2.0 ± 1.0 (n = 15; P < 0.001) after dynamic-clamp subtraction of 12.5, 25, 50, and 100% NaV1.7 conductance, respectively (figure 4, A and B). In the same cells, current threshold increment significantly increased from 3.9 ± 0.7 to 9.4 ± 1.6% (n = 13; P < 0.01), 17.1 ± 2.3% (n = 13; P < 0.001), and 32.8 ± 3.4% (n = 13; P < 0.001) in response to 12.5, 25, 50, and 100% reduction of NaV1.7 conductance. Since the NaV1.7 channel begins to activate at –55 to –50 mV (figure 1A), it is possible that subtraction of NaV1.7 conductance results in a reduced sodium current at subthreshold voltages, lowering the net current influx so that, to achieve AP threshold, the stimulus intensity would need to be increased to compensate for the lower sodium charge. Our measurements of NaV1.7 channel activity in a dynamic-clamped DRG neuron support this hypothesis. NaV1.7 began to activate at about –53 mV, reached 62% of its peak value at–32-mV threshold voltage, and reached peak amplitude at –19 mV (figure 4C). In the example shown, NaV1.7 sodium influx at subthreshold voltages comprises ~21% of the total sodium charge due to NaV1.7 channel activity occurring in the time interval between stimulus onset and AP undershoot (figure 4C). Subsequently, the channel inactivates quickly, and the resulting current is not present during the interspike interval provided membrane potential is not ramping back to the threshold level (figure 4D, but see figure 5C). The relatively slow channel repriming kinetics at membrane voltages close to resting membrane potential (RMP; figure 1C) still allow the channel to recover from inactivation during 10-Hz stimulation cycle without significant accumulation of inactivation (figure 4D, bottom). These data suggest that NaV1.7 channel activity at subthreshold membrane voltages drives the set point of current threshold for AP generation, thus regulating AP firing probability.

10310_006b_fig_004.jpg

Figure 4 Removal of NaV1.7 conductance raises current threshold for AP generation and reduces AP firing probability. (A) APs were evoked by 10-ms-long current pulses applied at 1.5× threshold amplitude in control (Ctrl; black) and after electronic subtraction of the incremental values of NaV1.7 conductance. Dynamic-clamp subtraction of NaV1.7 conductance (expressed as the incremental decrease over the endogenous NaV1.7 conductance) is noted on the y-axis. Scale bar: 200 ms. (B, top) Averages (n = 13) of current threshold change in response to the subtraction of respective proportion of endogenous NaV1.7 conductance. The solid line is a linear regression fit (r2 = 0.99) of the data. Statistical analysis on the top was performed between threshold increments obtained at 12.5% and at the respective percentage of conductance increment. (B, bottom) Averages (n = 15) of the number of APs evoked by the protocol described in A and plotted as a function of dynamically subtracted NaV1.7 conductance; the solid line is a linear regression fit (r2 = 0.98) of the data. **P < 0.01 and ***P < 0.001. (C) AP (top) evoked by a 10-ms-long current stimulus of 1.5× threshold intensity in control (black) and after dynamic-clamp subtraction of 12.5% of NaV1.7 conductance (blue) and the respective NaV1.7 model current (bottom). (D) APs (top) evoked by the protocol described above in control (black) and after dynamic-clamp subtraction (middle) of 12.5% of endogenous NaV1.7 conductance. The IV phase plot of the model NaV1.7 conductance dynamically subtracted during neuronal repetitive firing is shown on the bottom. Note that the positive-going (outward) dynamic-clamp current shown in C and D is flipped over 0 line to facilitate comparison of NaV1.7 current across the manuscript.

10310_006b_fig_005.jpg

Figure 5 Kinetic model of L858H NaV1.7 channel predicts a substantial enhancement of the persistent sodium current (INa) during repetitive firing of DRG neuron. (A) Normalized IV relationships obtained from WT (blue) and L858H (LH; red) NaV1.7 channel models. Respective traces of the INa = gmaxm3h(VmENa) model, where Vm is membrane voltage potential and ENa = 65 mV is sodium reversal potential, used to obtain the IV plot are presented on top. (B) Comparison of steady-state inactivation and steady-state activation (left), steady-state channel open probability (Po; middle), and activation time constant (τ; right) of the kinetic model of WT (blue) and L858H (red) NaV1.7 channel. (C, top) The trace shows repetitive firing of a spontaneously active small DRG neuron. The neuron spontaneously fired APs under i=o current-clamp conditions, i.e., with no additional injected current. A small –50-pA constant current was injected to stop AP firing. The trace presented in C, top, was recorded in response to the removal of the stabilizing –50-pA current. (C, bottom) Modeled WT (blue) and L858H (red) NaV1.7 currents obtained in response to the voltage command shaped in the form of the AP shown in C, top. Current-clamp recordings shown in C, top, were obtained from small DRG neuron in primary culture transfected by electroporation with WT NaV1.7 (Dib-Hajj et al. 2009). D: IV phase plots of WT (blue) and L858H (red) NaV1.7 currents presented in C, bottom. Both models were calculated in a 28-pF equipotential sphere of 1 µF/cm2 capacitance; conductance density was set to 0.029 S/cm2.

Model of NaV1.7 L858H Mutant Predicts an Enhanced Level of Channel Activity during Repetitive Firing of DRG Neuron

The L858H mutation in human NaV1.7 has been shown to be associated with a neuropathic pain syndrome, IEM (Yang et al. 2004). The mutation produces a hyperpolarizing shift in channel activation, slows channel deactivation, and causes an increase in amplitude of the NaV1.7 current in response to slow, small depolarizations (Cummins et al. 2004). When expressed in the native cell environment, transfection with mutant channels renders small DRG neurons hyperexcitable and is associated with depolarized RMP and reduced current threshold (Rush et al. 2006). To study the mechanism of NaV1.7 L858H-induced neuronal hyperexcitability in more detail, we performed quantitative evaluations of the effect of graded dynamic-clamp substitutions of L858H mutant conductance for WT NaV1.7 conductance on the electrical excitability of small DRG neurons. In the absence of empirical data, we assumed that the level of functional WT and L858H channels in the DRG neuron plasma membrane was the same.

First, we constructed a Hodgkin–Huxley (Hodgkin and Huxley 1952) model of the L858H channel based on the previously reported data (Cummins et al. 2004) by appropriately modifying our kinetic model of WT NaV1.7 channel. We adjusted activation rate constants to account for the –14-mV shift of L858H channel steady-state activation and for the altered mutant deactivation (figure 5B). The resulting kinetic model of L858H channel had a significantly enhanced window current (window current is defined as m3 × h at steady-state; figure 5B) and an apparent leftward shift of IV curve in a manner consistent with data reported by Cummins et al. (2004) (figure 5, A and B). Maximal steady-state channel open probability increased 17-fold from 9.5 × 10–5 at –32.3 mV to 1.6 × 10–3 at –51.5 mV (figure 5B); when measured at –63-mV RMP, it increased 255-fold from 3.8 × 10–6 to 9.7 × 10–3 (figure 5B). Here and thereafter, both WT and L858H models were calculated in a 28-pF equipotential sphere of 1 µF/cm2 capacitance; conductance density was set to 0.029 S/cm2. We further studied differences of WT and L858H channel behavior in response to a voltage command shaped in the form of membrane potential previously recorded from spontaneously active small DRG neuron (figure 5C, top). Both WT and L858H channels were active at subthreshold levels during ramplike membrane depolarizations between APs (figure 5C, left) as well as during APs (figure 5C, right). L858H model channel was already activated at a –62-mV initiation point producing –115-pA inward current, whereas WT channel produced only –0.5-pA current at this voltage. During the second cycle of AP firing, L858H channel began to activate at –70-mV undershoot (–2 pA) and reached –209 pA at –54 mV, whereas WT channel just began to activate (–2 pA) at –54 mV. At the –33.5-mV AP voltage threshold, WT current was –115 pA, whereas L858H current was –657 pA. L858H current peaked (–659 pA) at –32.7 mV, and WT current peaked (–320 pA) at –2.9 mV. Within the postpeak 0.5 ms (WT) and 1 ms (L858H), current declined sharply to about –20 pA at the 56.8-mV AP overshoot. This pattern of L858H channel activity resulted in a substantial enhancement of the sodium influx at subthreshold membrane voltages (WT, 4.1 × 10–13 C and L858H, 1.1 × 10–11 C) as well as during AP (WT, 2.8 × 10–13 C and L858H, 7.7 × 10–13 C). The differences in the pattern of WT and L858H channel activity remain unchanged throughout the repetitive cycle of AP firing (figure 5, C and D). Steady-state inactivation and inactivation kinetics were not affected by L858H mutation. Hence, both WT and L858H channels showed a similar level of use-dependent inactivation (figure 5, C and D).

The model predicts a substantial enhancement of NaV1.7 L858H mutant activity over a wide range of physiological membrane voltages during interspike intervals, at subthreshold levels of depolarization, and in the course of AP, thus increasing the sodium charge inflow.

Our kinetic model of the L858H mutant channel, consistent with the data published previously (Rush et al. 2006), predicted that expression of L858H mutant should enhance neuronal excitability. To test this hypothesis, we evaluated the effect of defined levels of L858H functional expression on electrical excitability of small DRG neurons using dynamic-clamp recording. A single-allele mutation of SCN9A, assuming equal efficiency of expression of WT and mutant channels, most probably results in 1:1 ratio of WT and L858H expressions in sensory neurons of the affected individual; however, the exact stoichiometry of WT-to-L858H expression is not known. We therefore used dynamic-clamp to assess the effect of substitutions of graded amounts of L858H current for WT current. In designing our experiments on rat DRG neurons, we first estimated the endogenous NaV1.7 conductance (see above for details) and then dynamically substituted graded amounts of endogenous conductance by equal amounts of L858H channel conductance. Essentially, we performed dynamic-clamp operation SR × gmax(L858H – WT).

We first tested the effect of WT-to-L858H substitution on current threshold. An AP was first evoked in a control DRG neuron without any current substitution by a 10-ms current pulse of threshold intensity (figure 6A). WT-to-L858H substitution was then implemented and resulted in a substantial inward current that developed at subthreshold voltages during stimulation and an apparent shortening of the delay for AP generation. The amplitude of the AP was not affected and was (from overshoot to undershoot) 113.9 mV in control vs. 114.8 (112.6) mV at a 25% (50%) SR [i.e., at 25% (50%) L858H]. At the same time, the maximal rate of AP rise became progressively smaller, being 114.2, 109.5, and 96.1 mV/ms in control and 25 and 50% SR, respectively (figure 6, A and B). We suggest that this deceleration of AP rate of rise could be the result of the additional inactivation of endogenous channels due to the depolarization of RMP described below; however, we cannot exclude the possibility that the reduction of net (L858H – WT) current seen at the respective voltage of AP maximal rise rate (20 mV; 527 pA at 25% SR and 385 pA at 50% SR) is a causative factor (figure 6B, right). Substitution of WT-to-L858H conductance in incremental amounts resulted in the reduction of AP current threshold in a linear manner (r2 = 0.96, figure 6C, top; r2 = 0.97, figure 6C, bottom). Current threshold was 668 ± 142 pA (control, n = 5) and was reduced to 598 ± 188, 457 ± 92, 328 ± 53, and 133 ± 20 pA (n = 5; P < 0.05) after 12.5, 25, 50, and 100% WT-to-L858H substitution, respectively (figure 6C). The reduction of current threshold was accompanied by enhancement of AP firing probability (figure 6D). A 10-Hz train of 10 current pulses applied at the threshold level evoked on average 1.3 ± 0.2 APs in control and 8.9 ± 0.5, 9.3 ± 0.6, 8.9 ± 1.0, and 8.4 ± 1.0 APs (n = 5; P < 0.01) after 12.5, 25, 50, and 100% WT-to-L858H substitution, respectively (figure 6E).

10310_006b_fig_006.jpg

Figure 6 L858H mutant lowers current threshold and enhances AP firing probability of DRG neuron. (A) AP (bottom, left) evoked by 10-ms-long stimulus (on top) of threshold intensity in control (black) and after implementation of 25% (green) and 50% (red) WT-to-L858H substitution ratio (SR); traces of the respective AP rate of change are shown in bottom, right. (B) Trajectory plots of AP rate of change (left) and the respective trajectory of stimulus (right) recorded at incremental levels of WT-to-L858H substitution; data are obtained from APs shown in A. (C) Current threshold (top) and current threshold change (bottom) obtained at different levels of WT-to-L858H conductance substitution in dynamically clamped DRG neuron. Solid line represents linear regression fit of the data (n = 5; top, r2 = 0.96; bottom, r2 = 0.97). Statistical analysis on the bottom was performed between threshold increments obtained at 12.5% and at the respective percentage of conductance increment. *P < 0.05. (D) APs evoked by a 10-Hz train of 10-ms-long current pulses of threshold intensity in control (black) and after incremental levels (the value is depicted on the left to the y-axis) of WT-to-L858H conductance substitution. Scale bar, 200 ms. (E) Averages (n = 5) of the number of APs evoked by the protocol presented in D at different levels of WT-to-L858H conductance substitution. **P < 0.01.

The present findings confirm earlier results (Rush et al. 2006) showing that the reduction of current threshold and the enhancement of AP firing probability are accompanied by RMP depolarization (figure 7C). At a mechanistic level, the present results add to the previous findings by showing that RMP depolarization is driven by the persistent window activity of L858H channel at membrane voltages close to RMP (figure 5B, middle). The effect of WT conductance subtraction in (L858H – WT) model on DRG neuron RMP was negligible since addition or subtraction of up to 100% of WT NaV1.7 conductance, performed in a set of additional experiments, did not significantly affect RMP of DRG neurons in the –65- to –62-mV range of recorded RMP (figure 7, A and B). The average RMP of DRG neurons was (mean ± SE, n = 6; P > 0.05): –64 ± 0.6 mV in control and –63.5 ± 0.8, –63.8 ± 0.7, –64.0 ± 0.5, –63.6 ± 0.6, –63.6 ± 0.6, –63.7 ± 0.7, –63.9 ± 0.8, and –64.0 ± 1.0 mV after addition of –100, –50, –25, –12.5, +12.5, +25, +50, and +100% WT conductance, respectively (figure 7, A and B). In contrast, WT-to-L858H substitution significantly depolarized RMP from –63.4 ± 0.6 mV in control to –62.3 ± 0.5 mV (n = 6; P > 0.05), –60.7 ± 0.8 mV (n = 6; P < 0.05), –58.4 ± 0.9 mV (n = 6; P < 0.01), and –56.6 ± 1.3 mV (n = 6; P < 0.01) after 12.5, 25, 50, and 100% WT-to-L858H conductance exchange, respectively (figure 7D).

10310_006b_fig_007.jpg

Figure 7 L858H mutant drives depolarization of resting membrane potential (RMP) of DRG neuron. (A) Dynamic-clamp recordings of RMP (bottom) and WT current (top) of DRG neuron in control and after addition of 100% WT NaV1.7 conductance. (B) Averages (n = 6) of the RMP (top) and the RMP changes (bottom) as a function of addition or subtraction of the incremental values of WT NaV1.7 conductance. (C and D) Data description is similar to A and B, but the RMP data were obtained as a response of WT-to-L858H substitution (WT conductance was dynamically subtracted while the equivalent amount of L858H conductance was added) at incremental levels ranging from 12.5 to 100% (n = 6). Statistical analysis on the bottom was performed between RMP changes obtained at 12.5% SR and at the respective percentage of conductance SRs. *P < 0.05 and **P < 0.01.

L858H Mutant Enhances DRG Neuron Excitability by Increasing Sodium Influx during Subthreshold Membrane Depolarizations and Interspike Intervals

Our model of the L858H mutant predicts an enhanced level of channel activity at subthreshold and suprathreshold voltages during repetitive firing of DRG neuron (see figure 5 and text above). The additional sodium charge inflow predicted by our results would be expected to be proexcitatory and should lead to enhanced neuronal excitability. We tested this hypothesis in dynamically clamped DRG neurons by assessing the response to a 10-Hz train of 10-ms depolarizing stimuli of two models: WT and (L858H – WT). We wanted to compare numerically the effect of these two models of channel activity on neuronal excitability side by side and first present data at 12.5% WT channel conductance addition and the respective 12.5% WT-to-L858H conductance exchange, since 12.5% SR is the minimal substitution studied for which AP firing is affected.

We found that L858H channels were already active at –63.4-mV RMP, producing current responsible for ~1-mV depolarizing shift of the RMP (–63.4 ± 0.6 mV in control and –62.3 ± 0.5 mV after 12.5% SR, n = 6, P > 0.05, but with a clear depolarizing trend of RMP), whereas WT current was not detectable above the noise, thus having no effect on RMP (–64 ± 0.6 mV in control and –63.6 ± 0.6 mV after addition of 12.5% WT conductance; figure 8, A–D). WT current began to activate (–3 pA) at –50.9 mV and reached its –216-pA maximum amplitude at 12.4 mV, a value of membrane voltages well above –40-mV AP threshold. In contrast, (L858H – WT) net current was already active (–21 pA) at –63-mV RMP and reached its maximum (–452 pA) at –30.6 mV. During interspike intervals, the net current resulting from (L858H – WT) substitution gradually increased from –1 pA at –69.5 mV to –20 pA at –64 mV, contributing to the development of slow, ramplike depolarization of the membrane potential (figure 8, D and E), whereas current produced by WT model was not detectable above the noise (figure 8, B and E). Significantly more net charge was carried by the (L858H – WT) compared with WT (in pA*ms/nA, we normalized current charge to the amplitude of native NaV1.7 current to account for cell-to-cell variability in the level of endogenous NaV1.7 current): 6.6 ± 1.2 (WT, n = 8) vs. 48 ± 10 (L858H, n = 5; P < 0.01) during subthreshold depolarization; and –0.2 ± 0.32 (WT, n = 8) compared with 58.2 ± 10.4 (L858H, n = 5; P < 0.01) during interspike intervals; whereas the charge during AP actually was smaller for mutant cells, 12.6 ± 6.3 (L858H, n = 5), compared with WT, 22 ± 3.5 (n = 8). These data indicate that the substantial increase of L858H channel activity at subthreshold membrane voltages results in a significant amplification of net sodium influx, subsequent depolarization of the membrane potential, reduction of current threshold, and ensuing enhancement of AP firing probability. These observations reveal enhancement of small DRG neuron excitability by substitution of as little as 12.5% of the NaV1.7 channels of the cell with L858H channel (i.e., with expression of the mutant channel at a density much lower than expected for a cell with 1 mutant allele) and demonstrate the powerful effect of the mutant channels on nociceptor excitability.

10310_006b_fig_008.jpg

Figure 8 L858H mutant augments DRG neuron excitability by enhancing sodium influx at subthreshold (subthresh) membrane voltages. (A) AP evoked by a 10-ms-long current pulse of threshold intensity in control (black) and after dynamically introduced +12.5% WT conductance (blue). APs are shown on the top (stimulation protocol is shown on top of APs), and the respective NaV1.7 current is presented in the bottom. (B) AP repetitive firing (top) evoked by a 10-Hz train of 10-ms-long current pulses at threshold intensity in control (black) and after 12.5% WT addition (blue); dynamic-clamp recording of the 12.5% WT current addition is shown on the bottom. (C and D) Same protocols as shown in A and B, but the comparison is made between data obtained in control (black) and after dynamic-clamp substitution of 12.5% WT to 12.5% L858H conductance. (E) I–V phase plots of dynamic-clamp recordings of repetitive AP firing in DRG neuron presented in B and D at +12.5% WT (blue) and at 12.5% WT to 12.5% L858H substitution (red). (F) Modeled sodium currents were recorded in dynamic-clamp mode and subsequently integrated over 3 different time intervals: (1) from stimulus onset to AP threshold (threshold is defined when 2nd differential of AP changes its sign); this interval extends from arrow 1 to 2 in A and C; (2) from threshold to undershoot; this interval extends between arrows 2 and 3 in A and C; and (3) from undershoot to the next stimulus onset. Sodium charge (pA*ms) was normalized to the peak value of the native NaV1.7 sodium current measured in each DRG neuron in voltage-clamp (see Materials and Methods). The data represent the model channel activity at +12.5% WT addition (n = 8, blue) and 12.5% WT-to-L858H substitution (n = 5, red). **P < 0.01.

Finally, as a model of nociceptors in patients carrying the L858H mutation, where single-allele mutation of SCN9A probably results in a ratio close to 1:1 of WT and L858H, we assessed the effect of a 50% SR of L858H channels. The model produced –37 ± 2 pA (n = 6) persistent current at rest (figure 9, A and B), which depolarized RMP on average by 5 mV from –63.4 ± 0.6 mV in control to –58.4 ± 0.9 mV (n = 6; P < 0.01) after 50% WT-to-L858H conductance exchange (figure 7D and figure 9, A and B). This current activated further at subthreshold membrane potentials in response to depolarizing current injection, reached –432 ± 112 (n = 6) peak at 2.7 ± 0.4 ms poststimulus before AP threshold, and then promptly declined to essentially zero level within the next 3.6 ± 0.7 ms (figure 9A). The net charge inflow due to the (L858H – WT) model activity during subthreshold depolarization and during AP was 0.8 ± 0.3 pC and 0.3 ± 0.2 pC (n = 6), respectively (figure 9C). During first interspike intervals, the net current resulting from 50% (L858H – WT) SR produced 2.9 ± 0.5 pC charge inflow, contributing to the development of slow, ramp-like depolarization of the membrane potential (figure 9, B and C). These observations of increased sodium influx underlying enhancement of small DRG neuron excitability following L858H substitution provide a mechanistic link between altered function of mutant channels and nociceptor hyperexcitability underlying the pain phenotype in patients carrying the NaV1.7 L858H mutation.

10310_006b_fig_009.jpg

Figure 9 A single-allele SCN9A mutation: L858H functional evaluation in small DRG neuron. (A) AP evoked by a current pulse of threshold intensity in control (black) and after dynamic-clamp 50% exchange of WT-to-L858H conductance (red). APs are shown on the top (stimulation protocol is shown on top of APs), and the respective NaV1.7 current differential is presented in the bottom. (B) AP repetitive firing (top) evoked by a 10-Hz train of current pulses (same as in A) at threshold intensity in control (black) and after 50% WT-to-L858H SR (blue); dynamic-clamp recording of the 50% (L858H – WT) current is shown on the bottom. (C) Dynamic-clamp (L858H – WT) model currents at 50% SR of endogenous NaV1.7 conductance were integrated over 3 different time intervals: (1) from stimulus onset to AP threshold (arrow 1 to 2 in A); (2) from threshold to undershoot (arrows 2 and 3 in A); and (3) from undershoot to the next stimulus onset. Sodium charge (pA*ms) was normalized to the peak value of the native NaV1.7 sodium current (n = 6). Kruskal-Wallis ANOVA nonparametric test for 3 populations was used to determine whether the samples come from different populations (P < 0.01).

Discussion

Our dynamic-clamp recording in native rat DRG neurons shows that increasing the NaV1.7 conductance density lowers threshold for a single AP and increases the number of APs fired in response to a train of depolarizing stimuli. Our data show a linear inverse relationship between functional NaV1.7 conductance and current threshold (the minimal stimulus capable of eliciting an AP). Consistent with the latter, we found a direct correlation between NaV1.7 conductance and AP firing probability. The relationship between NaV1.7 conductance and AP firing probability was also linear in the 0–25% range of additional NaV1.7 conductances. Saturation at 50–100% level was at least in part due to reaching maximal possible number of APs that could be evoked by our stimulation protocol.

It is generally accepted that TTX-S channels, including NaV1.7 channel, can function as a subthreshold sodium channel that amplifies membrane response to small depolarizing stimuli at subthreshold membrane voltages both in DRG neuron soma (Blair and Bean 2002; Choi and Waxman 2011; Kovalsky et al. 2009; Rush et al. 2007) and at the axon endings of primary sensory neurons (De Col et al. 2008; Pinto et al. 2008; Vasylyev and Waxman 2012). The functional impact of NaV1.7 channel activity on neuronal excitability and pain signal processing has been generally studied using an “all-or-none” paradigm using genetic knockout (Nassar et al. 2004) of NaV1.7 channel function. Functional contributions of NaV1.7 to DRG neuron excitability, including current and voltage thresholds for AP generation and AP repetitive firing, have also been studied using computer simulations of DRG neurons (Choi and Waxman 2011; Herzog et al. 2001; Kouranova et al. 2008; Kovalsky et al. 2009; Sheets et al. 2007). However, dynamic-clamp recording is superior to computer simulation for quantitative study of NaV1.7 channel function because it records the response of native neurons without making assumptions regarding which conductances to include in the computer model (Kemenes et al. 2011; Samu et al. 2012; Sharp et al. 1993). Our immediate and quite unexpected observation was that NaV1.7, when studied in the native DRG neuron environment, regulated the set point for AP current threshold in a remarkably linear manner: increasing or reducing NaV1.7 conductance by as little as 12.5% or as much as 100% produced a graded effect with a high degree of linearity over a 200% range of NaV1.7 conductances. We found that substitution of as little as 12.5% of channels with L858H produces hyperexcitability in DRG neurons. This observation predicts that expression of L858H produces hyperexcitability of DRG neurons even if the 50% reduction of current density seen after expression of L858H in HEK-293 cells (Cummins et al. 2004) applies to nociceptors.

Within the domain permitted by our stimulation protocol, NaV1.7 also regulated AP firing probability in a linear manner. This observation may be relevant to the pathophysiology of acquired pain since abnormal accumulations of NaV1.7 have been demonstrated at nerve endings within painful neuromas in rats (Persson et al. 2011) and humans (Black et al. 2008), and NaV1.7 levels and TTX-S current density have been shown to increase in DRG neurons in response to inflammation (Black et al. 2004) and in diabetic rats (Chattopadhyay et al. 2008, 2011). We would note, however, that although our previous patch-clamp analysis of small-diameter DRG neuron axons in vitro demonstrated a resting potential similar to that in DRG neuron somata, and sequential activation during AP clamp of TTX-S and TTX-resistant currents with characteristics attributed to NaV1.7 and NaV1.8 (Vasylyev and Waxman 2012) similar to that seen in DRG neuron somata (Blair and Bean 2002), we cannot exclude the possibility that the properties and functional role of NaV1.7 are not identical in DRG neuron somata vs. sensory axons and their terminals.

Dynamic-clamp requires an input of a kinetic model of the channel under study. Several kinetic models of TTX-S sodium channels, including NaV1.7 channel, have been recently proposed (Gurkiewicz et al. 2011; Herzog et al. 2001; Kovalsky et al. 2009; Sheets et al. 2007). These models rely on experimental data for TTX-S channels steady-state and kinetic properties with some degree of variability since these were obtained in different studies. This variability may be attributed, at least in part, to differences in recording solutions and in voltage protocols. TTX-S channel recordings were often performed using fluoride-based intracellular solution, which shifts the voltage dependence of NaV1.7 channel steady-state and kinetic parameters (Coste et al. 2004; Meadows et al. 2002; Saab et al. 2003; Sheets et al. 2007). At the same time, current-clamp recordings of AP firing in DRG neurons are commonly performed in a physiologically relevant chloride-based intracellular solution (Dib-Hajj et al. 2009; Estacion et al. 2011; Rush et al. 2006). It is reasonable to develop the kinetic model of NaV1.7 channel based on voltage-clamp data obtained under conditions similar to those where this model is utilized. Thus we developed a kinetic model of recombinant NaV1.7 channel based on voltage-clamp recordings obtained using intracellular solution essentially similar to that used for dynamic-clamp; the single difference between voltage-clamp and dynamic-clamp pipette solutions was that potassium chloride was replaced by cesium chloride on 1:1 basis (see Materials and Methods). We also used the same HBSS bath solution for voltage-clamp and dynamic-clamp recordings. Additionally, we performed a detailed analysis of NaV1.7 channel activation kinetics. Although an m3 model is generally accepted for sodium channel gating in squid giant axon (Hodgkin and Huxley 1952), mammalian skeletal muscles (Chanda and Bezanilla 2002), and mammalian sensory neurons (Herzog et al. 2001; Kostyuk et al. 1981; Ogata and Tatebayashi 1993; Sheets et al. 2007), an m2 model of TTX-S sodium channel activation has been suggested in mammalian central neurons (Baranauskas and Martina 2006). Consistent with previous studies of TTX-S sodium current (Herzog et al. 2001; Kostyuk et al. 1981; McCormick et al. 2007; Ogata and Tatebayashi 1993; Sheets et al. 2007), we found that recombinant NaV1.7 channel activated with an apparent delay, suggesting multiple transitions between closed states, in a sigmoidal manner with activation kinetics best described by m3 model.

The NaV1.7 L858H mutation was identified in a patient with hereditary primary erythromelalgia (Yang et al. 2004). Subsequent studies showed that the L858H mutation produces a –13.5-mV hyperpolarizing shift of NaV1.7 channel activation, slows channel deactivation, and enhances NaV1.7 current amplitude in response to slow, small depolarizations in a manner consistent with its erythromelalgia phenotype (Cummins et al. 2004). We modeled kinetics of the L858H mutant by appropriately adjusting WT channel activation rate constants to obtain the respective alterations in kinetics and in IV relationships similar to those reported for recombinant WT and L858H channels (Cummins et al. 2004). Our L858H model predicted enhancement of NaV1.7 channel activity at subthreshold membrane voltages. In response to a voltage command shaped in the form of membrane voltage of spontaneously firing DRG neuron, substitution of only 12.5% of the channels of the cell with the L858H model, compared with WT, resulted in a 27-fold increase of sodium influx at subthreshold membrane voltages and in a 3-fold sodium influx increase during the AP. When we evoked APs in dynamic-clamp using a 10-ms depolarizing stimulus, persistent L858H channel activity at subthreshold voltages produced >600% additional sodium influx during small depolarizations and resulted in the appearance of a significant sodium influx during interspike intervals (–0.2 ± 0.32, WT; compared with 58.2 ± 10.4, L858H). Such a substantial increase of the sodium influx is proexcitatory, and we were not surprised to observe a hyperexcitable neuronal phenotype due to L858H introduced by dynamic-clamp. Substitution of 50% WT-to-L858H conductance in our dynamic-clamp experiments produced persistent current that depolarized RMP on average by 5 mV and resulted in a reduction of current threshold on average by 51% without changing AP overshoot. This persistent current activated further at subthreshold membrane potentials in response to depolarizing current injection, reached its peak before AP threshold, and then promptly declined to essentially zero level within the next few milliseconds. The net charge inflow due to (L858H – WT) model was larger during subthreshold depolarization than during the AP and produced a significant charge inflow during interspike intervals driving a slow, ramplike depolarization of the membrane potential. On the basis of the present results, we attribute the L858H-induced RMP depolarization described by Rush et al. (2006) and seen in this study to the persistent current due to window channel activity. Steady-state open channel probability at –63-mV RMP increased 255-fold from 3.8 × 10–6 (WT) to 9.7 × 10–3 (L858H). Maximal conductance of endogenous NaV1.7 current (see Materials and Methods and Results) was 317 ± 68 nS (n = 6, WT experiments) and 354 ± 64 nS (n = 6, L858H experiments). Addition of 50% of the respective conductance resulted, assuming the calculated steady-state open probability, in 0.08-pA (WT) and 22-pA (L858H) persistent current at –63-mV RMP for dynamic-clamped neuron.

Our results obtained via dynamic-clamp at physiological levels of WT and mutant NaV1.7 conductance show, for the first time, that current threshold of small DRG neurons is regulated by NaV1.7 in a linear manner. Our observations also demonstrate that persistent activity of the L858H mutant channel in small DRG neurons amplifies sodium influx at subthreshold membrane voltages, so as to depolarize RMP and reduce current threshold for AP generation, thus producing hyperexcitability in nociceptive DRG neurons. Taken together, these findings establish a quantitative mechanistic link between the altered biophysical properties of a mutant NaV1.7 channel and nociceptor hyperexcitability underlying the pain phenotype in IEM.

Grants

This work was supported in part by grant from the Rehabilitation Research & Development Service and Medical Research Service, Department of Veterans Affairs to S. G. Waxman. The Center for Neuroscience and Regeneration Research is a collaboration of the Paralyzed Veterans of America with Yale University.

About the Authors

Dmytro V. Vasylyev, Department of Neurology and Center for Neuroscience and Regeneration Research, Yale University School of Medicine, New Haven, CT; and Rehabilitation Research Center, Veterans Affairs Connecticut Healthcare System, West Haven, CT

Chongyang Han, Department of Neurology and Center for Neuroscience and Regeneration Research, Yale University School of Medicine, New Haven, CT; and Rehabilitation Research Center, Veterans Affairs Connecticut Healthcare System, West Haven, CT

Peng Zhao, Department of Neurology and Center for Neuroscience and Regeneration Research, Yale University School of Medicine, New Haven, CT; and Rehabilitation Research Center, Veterans Affairs Connecticut Healthcare System, West Haven, CT

Sulayman Dib-Hajj, Department of Neurology and Center for Neuroscience and Regeneration Research, Yale University School of Medicine, New Haven, CT; and Rehabilitation Research Center, Veterans Affairs Connecticut Healthcare System, West Haven, CT

Stephen G. Waxman, Department of Neurology and Center for Neuroscience and Regeneration Research, Yale University School of Medicine, New Haven, CT; and Rehabilitation Research Center, Veterans Affairs Connecticut Healthcare System, West Haven, CT

Disclosures

No conflicts of interest, financial or otherwise, are declared by the author(s).

Author Contributions

D.V.V., S.D.-H., and S.G.W. conception and design of research; D.V.V., C.H., and P.Z. performed experiments; D.V.V. and C.H. analyzed data; D.V.V., C.H., S.D.-H., and S.G.W. interpreted results of experiments; D.V.V. and C.H. prepared figures; D.V.V. drafted manuscript; D.V.V., S.D.-H., and S.G.W. edited and revised manuscript; D.V.V., C.H., P.Z., S.D.-H., and S.G.W. approved final version of manuscript.

References

  1. Ahn HS, Vasylyev DV, Estacion M, Macala LJ, Shah P, Faber CG, Merkies IS, Dib-Hajj SD, Waxman SG. 2013. Differential effect of D623N variant and wild-type Na(v)1.7 sodium channels on resting potential and interspike membrane potential of dorsal root ganglion neurons. Brain Res 1529: 165–177.
  2. Baranauskas G, Martina M. 2006. Sodium currents activate without a Hodgkin- and-Huxley-type delay in central mammalian neurons. J Neurosci 26: 671–684.
  3. Black JA, Liu S, Tanaka M, Cummins TR, Waxman SG. 2004. Changes in the expression of tetrodotoxin-sensitive sodium channels within dorsal root ganglia neurons in inflammatory pain. Pain 108: 237–247.
  4. Black JA, Nikolajsen L, Kroner K, Jensen TS, Waxman SG. 2008. Multiple sodium channel isoforms and mitogen-activated protein kinases are present in painful human neuromas. Ann Neurol 64: 644–653.
  5. Blair NT, Bean BP. 2002. Roles of tetrodotoxin (TTX)-sensitive Na+ current, TTX-resistant Na+ current, and Ca2+ current in the action potentials of nociceptive sensory neurons. J Neurosci 22: 10277–10290.
  6. Catterall WA, Goldin AL, Waxman SG. 2005. International Union of Pharmacology. XLVII. Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol Rev 57: 397–409.
  7. Chanda B, Bezanilla F. 2002. Tracking voltage-dependent conformational changes in skeletal muscle sodium channel during activation. J Gen Physiol 120: 629–645.
  8. Chattopadhyay M, Mata M, Fink DJ. 2008. Continuous delta-opioid receptor activation reduces neuronal voltage-gated sodium channel (NaV1.7) levels through activation of protein kinase C in painful diabetic neuropathy. J Neurosci 28: 6652–6658.
  9. Chattopadhyay M, Mata M, Fink DJ. 2011. Vector-mediated release of GABA attenuates pain-related behaviors and reduces Na(V)1.7 in DRG neurons. Eur J Pain 15: 913–920.
  10. Choi JS, Waxman SG. 2011. Physiological interactions between NaV1.7 and NaV1.8 sodium channels: A computer simulation study. J Neurophysiol 106: 3173–3184.
  11. Coste B, Osorio N, Padilla F, Crest M, Delmas P. 2004. Gating and modulation of presumptive NaV1.9 channels in enteric and spinal sensory neurons. Mol Cell Neurosci 26: 123–134.
  12. Cummins TR, Dib-Hajj SD, Waxman SG. 2004. Electrophysiological properties of mutant NaV1.7 sodium channels in a painful inherited neuropathy. J Neurosci 24: 8232–8236.
  13. Cummins TR, Howe JR, Waxman SG. 1998. Slow closed-state inactivation: A novel mechanism underlying ramp currents in cells expressing the hNE/PN1 sodium channel. J Neurosci 18: 9607–9619.
  14. Cummins TR, Waxman SG. 1997. 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.
  15. De Col R, Messlinger K, Carr RW. 2008. Conduction velocity is regulated by sodium channel inactivation in unmyelinated axons innervating the rat cranial meninges. J Physiol 586: 1089–1103.
  16. Dib-Hajj SD, Choi JS, Macala LJ, Tyrrell L, Black JA, Cummins TR, Waxman SG. 2009. Transfection of rat or mouse neurons by biolistics or electroporation. Nat Protoc 4: 1118–1126.
  17. Dib-Hajj SD, Cummins TR, Black JA, Waxman SG. 2010. Sodium channels in normal and pathological pain. Annu Rev Neurosci 33: 325–347.
  18. Estacion M, Han C, Choi JS, Hoeijmakers JG, Lauria G, Drenth JP, Gerrits MM, Dib-Hajj SD, Faber CG, Merkies IS, Waxman SG. 2011. Intra- and interfamily phenotypic diversity in pain syndromes associated with a gain-of-function variant of NaV1.7. Mol Pain 7: 92.
  19. Gurkiewicz M, Korngreen A, Waxman SG, Lampert A. 2011. Kinetic modeling of NaV1.7 provides insight into erythromelalgia-associated F1449V mutation. J Neurophysiol 105: 1546–1557.
  20. Herzog RI, Cummins TR, Ghassemi F, Dib-Hajj SD, Waxman SG. 2003. Distinct repriming and closed-state inactivation kinetics of NaV1.6 and NaV1.7 sodium channels in mouse spinal sensory neurons. J Physiol 551: 741–750.
  21. Herzog RI, Cummins TR, Waxman SG. 2001. Persistent TTX-resistant Na+ current affects resting potential and response to depolarization in simulated spinal sensory neurons. J Neurophysiol 86: 1351–1364.
  22. Hille B. 1978. Ionic channels in excitable membranes: Current problems and biophysical approaches. Biophys J 22: 283–294.
  23. Hodgkin AL, Huxley AF. 1952. A quantitative description of membrane current and its application to conduction and excitation in nerve. J Physiol 117: 500–544.
  24. Kemenes I, Marra V, Crossley M, Samu D, Staras K, Kemenes G, Nowotny T. 2011. Dynamic clamp with StdpC software. Nat Protoc 6: 405–417.
  25. Kostyuk PG, Veselovsky NS, Tsyndrenko AY. 1981. Ionic currents in the somatic membrane of rat dorsal root ganglion neurons-I. Sodium currents. Neuroscience 6: 2423–2430.
  26. Kouranova EV, Strassle BW, Ring RH, Bowlby MR, Vasilyev DV. 2008. 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 153: 1008–1019.
  27. Kovalsky Y, Amir R, Devor M. 2009. Simulation in sensory neurons reveals a key role for delayed Na+ current in subthreshold oscillations and ectopic discharge: Implications for neuropathic pain. J Neurophysiol 102: 1430–1442.
  28. McCormick DA, Shu Y, Yu Y. 2007. Neurophysiology: Hodgkin and Huxley model—still standing? Nature 445: E1–2; discussion E2–3.
  29. Meadows LS, Chen YH, Powell AJ, Clare JJ, Ragsdale DS. 2002. Functional modulation of human brain NaV1.3 sodium channels, expressed in mammalian cells, by auxiliary beta 1, beta 2 and beta 3 subunits. Neuroscience 114: 745–753.
  30. Nassar MA, Stirling LC, Forlani G, Baker MD, Matthews EA, Dickenson AH, Wood JN. 2004. Nociceptor-specific gene deletion reveals a major role for NaV1.7 (PN1) in acute and inflammatory pain. Proc Natl Acad Sci USA 101: 12706–12711.
  31. Neher E. 1992. Correction for liquid junction potentials in patch clamp experiments. Methods Enzymol 207: 123–131.
  32. Ogata N, Tatebayashi H. 1993. Kinetic analysis of two types of Na+ channels in rat dorsal root ganglia. J Physiol 466: 9–37.
  33. Persson AK, Black JA, Gasser A, Cheng X, Fischer TZ, Waxman SG. 2010. Sodium-calcium exchanger and multiple sodium channel isoforms in intraepidermal nerve terminals. Mol Pain 6: 84.
  34. Persson AK, Gasser A, Black JA, Waxman SG. 2011. Nav1.7 accumulates and co-localizes with phosphorylated ERK1/2 within transected axons in early experimental neuromas. Exp Neurol 230: 273–279.
  35. Pinto V, Derkach VA, Safronov BV. 2008. Role of TTX-sensitive and TTX-resistant sodium channels in Aδ- and C-fiber conduction and synaptic transmission. J Neurophysiol 99: 617–628.
  36. Rush AM, Craner MJ, Kageyama T, Dib-Hajj SD, Waxman SG, Ranscht B. 2005. Contactin regulates the current density and axonal expression of tetro-dotoxin-resistant but not tetrodotoxin-sensitive sodium channels in DRG neurons. Eur J Neurosci 22: 39–49.
  37. Rush AM, Cummins TR, Waxman SG. 2007. Multiple sodium channels and their roles in electrogenesis within dorsal root ganglion neurons. J Physiol 579: 1–14.
  38. Rush AM, Dib-Hajj SD, Liu S, Cummins TR, Black JA, Waxman SG. 2006. A single sodium channel mutation produces hyper- or hypoexcitability in different types of neurons. Proc Natl Acad Sci USA 103: 8245–8250.
  39. Saab CY, Cummins TR, Waxman SG. 2003. GTP gamma S increases Nav1.8 current in small-diameter dorsal root ganglia neurons. Exp Brain Res 152: 415–419.
  40. Samu D, Marra V, Kemenes I, Crossley M, Kemenes G, Staras K, Nowotny T. 2012. Single electrode dynamic clamp with StdpC. J Neurosci Methods 211: 11–21.
  41. Sharp AA, O’Neil MB, Abbott LF, Marder E. 1993. Dynamic clamp: Computer-generated conductances in real neurons. J Neurophysiol 69: 992–995.
  42. Sheets PL, Jackson JO 2nd, Waxman SG, Dib-Hajj SD, Cummins TR. 2007. A NaV1.7 channel mutation associated with hereditary erythromelalgia contributes to neuronal hyperexcitability and displays reduced lidocaine sensitivity. J Physiol 581: 1019–1031.
  43. Toledo-Aral JJ, Moss BL, He ZJ, Koszowski AG, Whisenand T, Levinson SR, Wolf JJ, Silos-Santiago I, Halegoua S, Mandel G. 1997. Identification of PN1, a predominant voltage-dependent sodium channel expressed principally in peripheral neurons. Proc Natl Acad Sci USA 94: 1527–1532.
  44. Vasylyev DV, Waxman SG. 2012. Membrane properties and electrogenesis in the distal axons of small dorsal root ganglion neurons in vitro. J Neurophysiol 108: 729–740.
  45. Yang Y, Wang Y, Li S, Xu Z, Li H, Ma L, Fan J, et al. 2004. Mutations in SCN9A, encoding a sodium channel alpha subunit, in patients with primary erythermalgia. J Med Genet 41: 171–174.