Skip to main content
Journal of Neurophysiology logoLink to Journal of Neurophysiology
. 2015 Feb 4;113(7):2618–2634. doi: 10.1152/jn.00717.2014

Experimental and computational evidence for an essential role of NaV1.6 in spike initiation at stretch-sensitive colorectal afferent endings

Bin Feng 1,, Yi Zhu 1, Jun-Ho La 1, Zachary P Wills 2, G F Gebhart 1
PMCID: PMC4416596  PMID: 25652923

Abstract

Stretch-sensitive afferents comprise ∼33% of the pelvic nerve innervation of mouse colorectum, which are activated by colorectal distension and encode visceral nociception. Stretch-sensitive colorectal afferent endings respond tonically to stepped or ramped colorectal stretch, whereas dissociated colorectal dorsal root ganglion neurons generally fail to spike repetitively upon stepped current stimulation. The present study investigated this difference in the neural encoding characteristics between the soma and afferent ending using pharmacological approaches in an in vitro mouse colon-nerve preparation and complementary computational simulations. Immunohistological staining and Western blots revealed the presence of voltage-gated sodium channel (NaV) 1.6 and NaV1.7 at sensory neuronal endings in mouse colorectal tissue. Responses of stretch-sensitive colorectal afferent endings were significantly reduced by targeting NaV1.6 using selective antagonists (μ-conotoxin GIIIa and μ-conotoxin PIIIa) or tetrodotoxin. In contrast, neither selective NaV1.8 (A803467) nor NaV1.7 (ProTX-II) antagonists attenuated afferent responses to stretch. Computational simulation of a colorectal afferent ending that incorporated independent Markov models for NaV1.6 and NaV1.7, respectively, recapitulated the experimental findings, suggesting a necessary role for NaV1.6 in encoding tonic spiking by stretch-sensitive afferents. In addition, computational simulation of a dorsal root ganglion soma showed that, by adding a NaV1.6 conductance, a single-spiking neuron was converted into a tonic spiking one. These results suggest a mechanism/channel to explain the difference in neural encoding characteristics between afferent somata and sensory endings, likely caused by differential expression of ion channels (e.g., NaV1.6) at different parts of the neuron.

Keywords: computational model, sodium channel, single fiber, colon, visceral pain


irritable bowel syndrome patients typically suffer from persistent pain and organ hypersensitivity, manifesting enhanced responses and reduced thresholds to mechanical distension of the distal colorectum (Naliboff et al. 1997). Understanding the afferent innervation of the colorectum is important, as targeting colorectal afferents has proven to be effective in alleviating pain and hypersensitivity in irritable bowel syndrome patients, e.g., intrarectal instillation of local anesthetics (Verne et al. 2003, 2005) and oral intake of the guanylate cyclase-C agonist linaclotide (Busby et al. 2013; Chey et al. 2012; Rao et al. 2012).

The colorectum in the mouse is innervated by lumbar splanchnic and pelvic nerve (PN) pathways (Brierley et al. 2004; Feng and Gebhart 2011). Using an in vitro colorectum-nerve preparation and an unbiased electric search strategy, we categorized colorectal afferents into five mechanosensitive classes [serosal, mucosal, muscular, muscular-mucosal (M/M), mesenteric] and one mechanically-insensitive class (Feng et al. 2012a). Among these classes, only muscular and M/M afferents tonically encode circumferential stretch of the colorectum (i.e., are stretch-sensitive) and subserve the encoding of nociceptive colorectal distension (Feng et al. 2010, 2013).

Patch-clamp recordings reveal that colorectal dorsal root ganglion (c-DRG) somata have small diameters (<25–30 μm), tetrodotoxin (TTX)-resistant inward sodium currents, a significant inflection/hump in the repolarization phase of the action potential (AP) (Beyak et al. 2004), and are immuno-positive for one or more neurochemical markers, including transient receptor potential vanilloid 1 (up to 85%), calcitonin gene-related peptide (up to 80%), and isolectin B4 (up to 20%) (Christianson et al. 2006; La et al. 2011). Consistent with findings from c-DRG somata, almost all mechanosensitive colorectal afferents in mice are unmyelinated C fibers with conduction velocities ≤1 m/s (Feng and Gebhart 2011).

Our laboratory's previous single-fiber studies suggest that a significant proportion (∼33%) of colorectal afferent endings in the PN innervation are stretch-sensitive and respond tonically to ramped circumferential colorectal stretch (0 to 170 mN in 34 s) (Feng and Gebhart 2011; Feng et al. 2012b, 2012c) and could be categorized as class-1 tonic-spiking neurons according to Hodgkin's scheme (Hodgkin 1948). In contrast, most if not all lumbosacral c-DRG somata (corresponding to the PN innervation) fail to spike repetitively, typically spiking only at the onset of stepped current stimulation (e.g., Fig. 2C in Shinoda et al. 2010), and can be categorized as class-3 single-spiking neurons. This discrepancy in spiking patterns between colorectal afferent somata and their sensory endings is likely contributed to by different compositions and/or densities of voltage-gated channels in their respective membranes. In small-diameter, presumptive nociceptive DRG neurons, most inward Na current is through voltage-gated sodium channel (NaV) 1.7 and NaV1.8 channels, whereas a NaV1.6 current is less common and found mainly in larger-diameter DRG neurons (Cummins et al. 2005). However, NaV1.6 contributes to the tonic firing of neurons due to its low activation threshold, rapid repriming, and large persistent and resurgent current (Rush et al. 2005). We hypothesized that NaV1.6 contributes significantly to the tonic firing of stretch-sensitive colorectal afferent endings and tested this hypothesis using combined single-fiber electrophysiology and computational simulation.

Fig. 2.

Fig. 2.

Voltage-gated sodium channel (NaV) 1.6 and NaV1.7 immunohistochemistry and Western blotting. A: a proportion of mouse L6 DRG neurons retrogradely labeled with Fast Blue (FB) from the distal colon express NaV1.6 and NaV1.7 (arrows). Not all FB-labeled neurons expressed NaV1.6 and NaV1.7 (arrowheads). Scale bars = 50 μm. B: NaV1.6 and NaV1.7 were colocalized with yellow fluorescent protein (YFP; arrows) expressed in sensory neurons (driven by SNS-Cre). MP, myenteric plexus; CM, circular muscle; M, mucosa. Arrows indicate colocalization of YFP-expressing nerve fibers with either NaV1.6- or NaV1.7-immunoreactivity. Many segments of YFP-expressing nerve fibers do not have NaV1.6- or NaV1.7-immunoreactivity (arrowheads). Scale bars = 10 μm. C: Western-blot detection of NaV1.6 and NaV1.7 in protein extracts from pelvic nerve (PN), DRG and colon. The protein quantity of the two channels was below detection sensitivity in colon due to the low content of neuronal proteins in colon tissue homogenates [note the small quantity of neuron-specific protein gene product (PGP) 9.5].

METHODS

Unless specified, experiments were conducted in male C57BL/6Tac mice, 6–8 wk old, 20–30 g (Taconic, Germantown, NJ), and approved by the University of Pittsburgh Institutional Animal Care and Use Committee.

Immunohistochemistry

As previously described (Feng et al. 2012c), mice that express yellow fluorescent protein (YFP) in sensory neurons driven by SNS-Cre (a gift from Dr. Brian M. Davis, University of Pittsburgh) were euthanized via CO2 inhalation. The L6 DRG and distal colorectum were harvested and fixed with 4% paraformaldehyde in 0.16 M phosphate buffer containing 14% picric acid (Sigma-Aldrich). After cryoprotection in 20% sucrose, fixed tissue was embedded in OCT compound (Sakura Finetek, Tokyo, Japan), frozen, and sectioned at 20 μm for DRGs and at 70 μm for colorectum. Tissue sections were incubated with antibodies against either NaV1.6 (1:1,000, Alomone, ASC-009, lot no. AN1750) or NaV1.7 (1:1,000, Alomone, ACS-008, lot no. AN1125) and co-stained with goat antibodies against YFP (1:1,000, Abcam, Cambridge, MA, AB5450, lot no. GR136040-1). The sections were further stained with Cy3-conjugated anti-rabbit IgG (1:200, Jackson Immunoresearch, 711-227-003, lot no. 95161) and Alexa Fluor 488-conjugated anti-goat IgG (1:200, Molecular Probes, Eugene, OR, A11055, lot no. 989791). Confocal microscopy was carried out with a Nikon A1R point scanning microscope with either a 0.75 numerical aperture ×20 objective, or 1.4 numerical aperture ×60 objective. Pinhole size was limited to 1 airy unit to maximize confocality of all images. Twelve-bit images were captured with limited laser and detector power to ensure fluorophore signals were not saturated. Z-stack images were captured in 1 μm (×20) or 0.5 μm (×60) increments. Three-dimensional reconstructions were rendered using Nikon elements software (version AR 4.13.04).

To localize the two sodium channels in DRG neurons innervating the colorectum via PNs, a retrograde tracer Fast Blue (1% in sterile saline, EMS-Chemie, Gross-Umstadt, Germany) was injected into the distal colon wall, as previously described (La et al. 2012). The L6 DRG were harvested 14 days later and processed for immunohistochemistry as above. Immunostained tissue sections were photographed using a microscope-mounted digital camera (DFC340FX; Leica).

Western Blot

As previously described (Zhu et al. 2012), tissue samples of L6 DRG, PN, and colon were individually homogenized with a Teflon tube and mortar for less than 10 strokes in ice-cold radioimmunoprecipitation assay buffer containing protease inhibitors (Sigma-Aldrich). Protein concentration was determined using a bicinchoninic acid protein assay kit (Thermo Scientific, Rockford, IL). The lysates were then mixed with Laemmli buffer (×6) and boiled for 5 min before loading. Samples (100 μg/lane) were then loaded and separated on a 5% SDS-PAGE gel and transferred to a nitrocellulose membrane. Membranes were blocked with 5% skim milk for 1 h at room temperature, then incubated with primary antibody at 4°C overnight (NaV1.6 or NaV1.7 both at 1:200, Alomone), and diluted with 5% milk/Tris-buffered saline with Tween 20 (TBST, Sigma-Aldrich). The blots were washed and then incubated with peroxidase-conjugated secondary antibody (1:2,000 in 5% milk/TBST, Jackson ImmunoResearch) for 1 h at room temperature. An ECL kit (Amersham Biosciences, Piscataway, NJ) was used for detection of immunoreactivity, and image of the blots was then taken with an LAS3000 imager (Fujifilm, Japan).

In Vitro Mouse Colon-PN Preparation

As detailed previously (Feng et al. 2013), mice were euthanized via CO2 inhalation followed by exsanguination after the right atrium was perforated. The distal colorectum with attached PN was dissected and transferred to ice-cold Kreb's solution bubbled with carbogen (95% O2, 5% CO2). The colorectum was opened longitudinally, pinned flat either mucosal or serosal side up, depending on the experiment objective in a tissue chamber, and the PN extended into an adjacent recording chamber filled with paraffin oil. The tissue chamber was superfused with a modified Krebs solution (in mM: 117.9 NaCl, 4.7 KCl, 25 NaHCO3, 1.3 NaH2PO4, 1.2 MgSO4·7H2O, 2.5 CaCl2, 11.1 d-glucose, 2 butyrate, and 20 acetate) at a temperature of ∼30–32°C to which nifedipine (4 μM) and indomethacin (3 μM) were added. The PN was teased into fine bundles (∼10 μm thickness) for single-fiber recording.

As previously described (Feng and Gebhart 2011), mechanosensitive colorectal afferents were classified as serosal, muscular, mucosal or M/M based upon responses to probing with calibrated nylon monofilaments (0.4, 1, and 1.4 g force), mucosal stroking (10 mg force) and circumferential stretch. Muscular and M/M afferents both respond to stretch, and their stimulus-response functions (SRFs) to circumferential stretch were generated using a servo-controlled force actuator (Aurora Scientific, Aurora, Ontario, Canada). Custom-made claws (∼1-mm interval) were inserted along the anti-mesenteric edge of the colorectum to permit homogeneous, circumferential stretch by a slow ramped force (0 to 170 mN at 5 mN/s) corresponding to intraluminal pressures of 0–45 mmHg (Feng et al. 2010).

Application of Chemicals to Afferent Endings

The following compounds were applied directly to afferent endings in the colorectum, as previously described (Feng and Gebhart 2011; Feng et al. 2013): TTX, the NaV1.8 antagonist A803467 (A803), the NaV1.7 antagonist ProTx-II (PTX), and NaV1.6 antagonists μ-conotoxin GIIIa (mCtxG) and μ-conotoxin PIIIa (mCtxP). After establishing a baseline (control) SRF, the receptive ending was isolated (4 × 4 mm2 ×10 mm high tubing) and the Krebs solution removed and replaced by 150 μl of TTX, A803, PTX, mCtxG or mCtxP for 5 min. The tubing was then removed, and a SRF acquired immediately afterwards (experiment). After reexposing the ending to Krebs solution for 15 min, a third SRF (wash) was acquired to conclude the protocol. Generally, each fiber was exposed to two to three protocols using different drugs or concentrations with at least 15 min wash between protocols. Afferent responses after exposure to A803 did not recover after wash, and those fibers were not tested with other drugs. To avoid direct application of drugs onto the PN fibrils, afferents with receptive fields within 5 mm of the nerve entry point into the colorectum were not studied pharmacologically when the colorectum was pinned serosal side up.

TTX was dissolved in Krebs solution at 3 mM and prepared in aliquots of 3.3 μl. A803 was dissolved in DMSO to 3 mM and prepared in nitrogen-filled aliquots of 3.3 μl (to avoid oxidation); DMSO (<0.3%) alone in Krebs solution has no significant effect on afferent responses to stretch (Feng et al. 2013; Kiyatkin et al. 2013). PTX was dissolved in 80% acetonitrile to 0.3 mM and prepared in aliquots of 5 μl; acetonitrile (0.8%) alone in Krebs solution has no significant effect on afferent responses to stretch (data not shown). mCtxG and mCtxP were dissolved in Krebs solution at 0.5 mM and 0.1 mM, respectively, and prepared in aliquots of 42 μl. All aliquots were frozen, stored at −20°C and diluted on the day of an experiment to final concentrations in freshly oxygenated Krebs solution (TTX, PTX, mCtxG and mCtxP) or Krebs solution with pH adjusted to 7.4 by hydrochloric acid (A803, to avoid oxidation). TTX, PTX and A803 were purchased from Tocris (Bristol, UK), mCtxG from VWR (Radnor, PA) and mCtxP from Alomone. All other chemicals were purchased from Sigma-Aldrich.

Circumferential Colorectal Stretch

Circumferential colorectal stretch was quantified as a stretch ratio (λ) by measuring colorectal circumference during ramped stretch in vitro. Colorectal deformation was recorded through a stereo dissection microscope using a charge-coupled device camera; images were extracted every 2 s during stretch, and circumference was measured using ImageJ (v1.44p, National Institutes of Health).

Computational Simulation

Model geometry.

To facilitate comparisons, electrophysiological properties of colorectal afferent endings and DRG somata were simulated separately in the NEURON simulation environment (Carnevale and Hines 2005). The fundamental morphological and electrical features of a mouse colorectal afferent ending was represented by a multicompartmental model, as depicted in Fig. 1A. The nerve terminal model presented here is designed to emulate AP encoding, assuming a transducer terminal region contiguous with a single spike-initiation zone (siz) similar to those developed to simulate the axon initial segment (AIS) of neurons in the central nervous systems (e.g., Hu et al. 2009). The model consists of a transducer zone (trsd), where a generator potential is produced by a depolarizing current from mechanosensitive channels (ms channels), a siz where AP spikes are evoked by the generator potential, and a middle section (mid) in between in which Na+ and K+ channel densities gradually increase from the trsd side to siz side (to simulate the gradual change in ion channel densities); the passive compartment (pas) distal to siz provides space for axial diffusion of intracellular Na+ and K+ ions. All four compartments are cylinders 0.8 μm in diameter. To achieve spatial and temporal accuracy in simulation, the compartments were further divided into a total of 23 segments (10 in trsd, 5 each in siz and mid, and 3 in pas), so each segment length was less than 1/50th of the electrotonic length constant (570 μm in the model) (Carnevale and Hines 2005). In contrast, dissociated DRG somata free of attached axons were simulated as a single-segment model depicted in Fig. 1B and assigned a diameter of the average colorectal DRG neuron (Φ 23 μm) (Beyak et al. 2004).

Fig. 1.

Fig. 1.

Schematics for the computational simulation of action potential generation in colorectal afferent endings and dissociated dorsal root ganglion (DRG) somata. A: the colorectal afferent ending is simulated by a four-compartment cylinder consisting of a transducer (trsd), a spike initiation zone (siz), a transition zone in the middle (mid), and a passive conducting zone (pas) proximal to siz. B: The soma is simulated as a one-segment membrane model. Iclamp and Vclamp, clamp current and voltage, respectively. C: the afferent ending model simulates encoding of mechanical stretch by inclusion of mechanosensitive (ms) channels in the trsd zone that are gated by the membrane tension τ. A lumped parametric model is used to translate bulk colorectal deformation (stretch force F, stretch ratio λ, and their first derivatives, Ḟ, λ̇) to the membrane tension τ at the afferent ending. C also includes a representative ms current in response to a stepped colorectal stretch at 100 mN. ΔW, unit width of the simulated neural membrane patch; ΔL, unit length of the simulated neural membrane patch.

Voltage-gated ion channels and pump.

The model incorporates four different Na+ conductances representing NaV1.6, NaV1.7, NaV1.8, and NaV1.9 and three K+ conductances, simulating the fast inactivating A-type current (KA), slowly inactivating A-type current (KD), and sustained current (KS). NaV1.6 and NaV1.7 channels were represented by Markov models with multiple gating states to capture their unique and contrasting gating features (e.g., rapid vs. slow repriming and incomplete vs. complete inactivation). The other channels were modeled by Hodgkin-Huxley formulations. Na+-K+-ATPase was simulated as a voltage- and intracellular Na+ concentration ([Na+]i)-dependent outward current with a 3:2 transport ratio between Na+ and K+ ions. [Na+]i and intracellular K+ concentration ([K+]i) are dynamically influenced by ion flow across the membrane via channels, pumps, and leak conductances, as well as by passive axial diffusion, assuming a diffusion coefficient of 0.6 μm2/ms (Fleidervish et al. 2010; Rugiero et al. 2010). Na+ and K+ reversal potentials were derived from the ion concentrations across the membrane.

NAV1.6.

NaV1.6 is simulated by a Markov-type model (see Fig. 6A) adopted from Khaliq et al. (2003) (see also Raman and Bean 2001). NaV1.6 current is determined by the open probability <O> and membrane potential V: INa6 = Na6 <O> (VENa), where INa6 is NaV1.6 current, Na6 is maximum NaV1.6 conductance, and ENa is sodium reversal potential. Rate coefficients (ms) that are membrane voltage-dependent (mV) are adopted from Khaliq et al. (2003), except that ζ was adjusted to 0.6 exp(−V/25) to accommodate to the low firing rate of colorectal afferent neurons.

Fig. 6.

Fig. 6.

Effect of the selective NaV1.8 blocker A803467 (A803) on stretch-sensitive colorectal afferent endings. When applied from the serosal side, afferent responses to ramped stretch were unaffected by either 1 μM (A) or 3 μM A803 (C). B: mucosal application of 3 μM A803 did not affect responses to stretch, whereas 10 μM A803 slightly, but significantly, reduced responses (F2,20 = 6.0, P = 0.009, post hoc comparison vs. control, P = 0.01). D: response thresholds (insets in AD) were unaffected by A803. E and F: normalized responses to stretch (total spike number) are plotted for A803 applied from the serosal and mucosal side, respectively. There were no significant differences between control, A803, and wash groups. *P < 0.05.

NAV1.7.

NaV1.7 is simulated by a Markov-type model (see Fig. 6A) adopted from Gurkiewicz et al. (2011). NaV1.7 current is determined by the open probability <O> and membrane potential V: INa7 = Na7 <O> (VENa), where INa7 is NaV1.7 current and Na7 is maximum NaV1.7 conductance. All model parameters were adopted from Gurkiewicz et al. (2011).

NAV1.8.

NaV1.8 is simulated by a Hodgkin-Huxley-type model adopted from Baker (2005).

INa8=g¯Na8m3h(VENa)
m˙=mmτm
h˙=hhτh
αm=3.83/[1+exp(V+2.5811.47)]
βm=6.894/[1+exp(V+61.219.8)]
m=αmαm+βm
τm=1.0αm+βm
αh=0.013536exp(V+10546.33)
βh=0.61714/[1+exp(V21.811.998)]
h=αhαh+βh
τh=1.0αh+βh

where INa8 is NaV1.8 current; Na8 is maximum NaV1.8 conductance; ∞ is infinity; m and h are Hodgkin-Huxley typed voltage-gating parameters for sodium channels; and are first derivatives of m and h, respectively; τ is neural membrane tension; and α and β are rate constants of the ms channel as functions of τ.

NAV1.9.

NaV1.9 is simulated by a Hodgkin-Huxley-type model adopted from Baker (2005).

INa9=g¯Na9mh(VENa)
m˙=mmτm
h˙=hhτh
αm=1.548/[1+exp(V11.0114.871)]
βm=8.685/[1+exp(V+112.422.9)]
m=αmαm+βm
τm=1.0αm+βm
αh=0.2574/[1+exp(V+63.2643.7193)]
βh=0.53984/[1+exp(V+0.278539.0933)]
h=αhαh+βh
τh=1.0αh+βh

where INa9 is NaV1.9 current; and Na8 is maximum NaV1.9 conductance.

KA.

KA is simulated by a Hodgkin-Huxley-type model adopted from Schild et al. (1994) with minor adjustments of parameters.

IKA=g¯KAp3q(VEK)
p˙=ppτp
q˙=qqτq
p=1.0/[1.0+exp(V+15.7910.0)]
τp=5.0exp[0.0222(V+65)2]+1.5
q=1.0/[1.0+exp(V+587.0)]
τq=45.0exp[0.00352(V+10)2]+10.5

where IKA is KA voltage-gated potassium current; KA is maximum KA conductance; p and q are Hodgkin-Huxley typed voltage-gating parameters for KA current; and are first derivative of p and q, respectively; and EK is potassium reversal potential.

KD.

KD is a Hodgkin-Huxley-type model adopted from Schild et al. (1994) with minor adjustments of parameters.

IKD=g¯KDx3y(VEK)x˙=xxτx
y˙=yyτyx=1.0/[1.0+exp(V+14.5915.0)]
τx=5.0exp[0.0222(V+65)2]+3.5
y=1.0/[1.0+exp(V+487.0)]
τy=1,800

where IKD is KD voltage-gated potassium current; KD is maximum KD conductance; x and y are Hodgkin-Huxley typed voltage-gating parameters for KD current; and and are first derivatives of x and y, respectively.

KS.

KS is a Hodgkin-Huxley-type model adopted from Schild et al. (1994) with minor adjustment of parameters.

IKS=g¯KSn4(VEK)
n˙=nnτn
n=αnαn+βn
τn=1.0αn+βn
αn=0.1(V+5212.5)/[1.0exp(V+5212.5)]
βn=0.125exp(V+6080)

where where IKS is KS voltage-gated potassium current; KS is maximum KS conductance; n is Hodgkin-Huxley typed voltage-gating parameters for KS current; and is first derivative of n.

NA+-K+-ATPASE.

Na+-K+-ATPase is adapted from Bondarenko et al. (2004) with parameters slightly adjusted.

INaK=I̅NaKfNaK11+(Km,Nai/[Na+]i)4[K+]o[K+]o+Km,Ko
INaK=INa,p+IK,p
INa,p=3INaK
IK,p=2INaK
fNaK=11+0.1245exp(0.1VFRT)+0.0365σexp(VFRT)
σ=17[exp([Na+]o67.3)1]
Km,Nai=13mM
Km,Ko=1.5mM

where INaK is total current from Na+-K+-ATPase activities; ĪNaK is maximum INaK; fNaK, Km,Nai, and Km,Ko are parameters for the Na+-K+-ATPase current; [K+]o is extracellular K+ concentration; INa,p is sodium current from Na+-K+-ATPase activities; IK,p is potassium current from Na+-K+-ATPase activities; V is membrane voltage; F is Faraday constant; R is universal gas constant; T is absolute temperature; [Na]o is extracellular Na concentration; and σ is a function of [Na+]o.

BACKGROUND CURRENT.
IB=g̅B(VEB)
IB=INa,B+IK,B
INa,B=g̅BEBEKENaEK(VENa)
IK,B=g̅BENaEBENaEK(VEK)

where IB is background current; B is maximum conductance of background; INa,B is background current contributed by sodium; IK,B is background current contributed by potassium; and EB is background current reversal potential.

The summary of the ion channels and Na+-K+-ATPase included in modeling is listed in Table 1.

Table 1.

Summary of voltage-gated ion channels and pump

Channel/Pump Model Type Reference
NaV1.6 13-state Markov Raman and Bean 2001
NaV1.7 6-state Markov Gurkiewicz et al. 2011
NaV1.8 Hodgkin-Huxley Baker 2005
NaV1.9 Hodgkin-Huxley Baker 2005
KA Hodgkin-Huxley Schild et al. 1994
KD Hodgkin-Huxley Schild et al. 1994
KS Hodgkin-Huxley Schild et al. 1994
Na+-K+-ATPase [Na+]i and Vm dependent Bondarenko et al. 2004

NaV, voltage-gated sodium channel; KA, fast inactivating A-type current; KD, slowly inactivating A-type current; KS, sustained current; [Na+]i, intracellular Na+ concentration; Vm, membrane potential.

Mechanosensitive ion channels.

To simulate the gating of a ms ion channel, we used a two-state model that includes an open state (O) and a closed state (C), the rates of transition between which are α and β, two exponential functions of membrane tension τ at the afferent ending:

ObaC
α=1Aexp(ττ02S)

where S is a parameter determining α and β.

β=1Aexp(ττ02S)

Assume that the fraction of ms channels in the open state is denoted by p, and we have:

p˙=ppTp

in which,

p=αα+β

and

Tp=1α+β

where is first derivative of p, and Tp is exponential decay time constant for p.

p, the open probability of the ms channel at steady state, follows Boltzmann's equation, consistent with previous theoretical and experimental studies on ms channels (Hao and Delmas 2010; Haselwandter and Phillips 2013; Wiggins and Phillips 2005).

p=11+exp(ττ0S)

To allow calculation of τ from bulk colorectal deformation (circumferential stretch force F, stretch ratio λ, and their derivatives Ḟ and λ̇), the passive mechanical properties of the colorectal wall tissue was simulated by a lumped parametric model consisting of two springs and one dashpot (Fig. 1C), which leads to the following equations:

{x2=(λ1)ΔLτΔw=k2(x2x1)τΔw=k1x1+c1x˙1

where x is Hodgkin-Huxley typed voltage-gating parameter for KD current; ẋ is first derivative of x; ΔL is unit length of the simulated neural membrane patch; Δw is unit width of the simulated neural membrane patch; k1 and k2 are two linear spring components in the lumped parametric model that transfers bulk colorectal deformation into neural membrane tension; and c1 is one linear dashpot component in the lumped parametric model.

Assuming:

Tτ=c1k1+k2
m1=k1k2ΔL(k1+k2)Δw
m2=k2c1ΔL(k1+k2)Δw

Then:

τ˙=m1(λ1)+m2λ˙τTτ

where Tτ is the exponential decay time constant for τ, and τ̇ is first derivative of τ.

The ms channel conductance (gM) is divided into conductance for Na+ (gNa,M) and K+ ions (gK,M):

gM=g̅Mp=gNa,M+gK,M

Assuming the conductance for Na+ and K+ are proportional to their respective driving forces, i.e., (ENaV) and (VEK), then we have:

INa,M=g¯MpENaVENaEK(VENa)
IK,M=g̅MpVEKENaEK(VEK)
IM=INa,M+IK,M

where ḡM is maximum conductance of ms current; INa,M is ms current contributed by sodium; and IK,M is ms current contributed by potassium.

Parameters for the ms channel in the study are as follows: parameter determining α and β (A), 10 ms; parameter determining α and β (τ0), 4.45 mN/m; parameter determining α and β (S), 2.07 mN/m; exponential decay time constant for τ (Tτ), 1,000 ms; parameter determining τ (m1), 13.6 mN/m; parameter determining τ (m2), 16,900 (mN·ms)/m.

The ms current evoked by a stepped colorectal stretch is plotted in Fig. 1C, which recapitulated features of a typical slowly adapting ms current observed experimentally in DRG neurons (Rugiero et al. 2010).

Passive properties and initial conditions.

The passive electrical properties, Cm (specific membrane capacitance of neural endings and somata), Rm (membrane resistivity of neural endings and somata), and Ri (axial resistivity of neural endings) in both models were set to 1 μF/cm2, 10,000 Ω/cm2, and 123 Ω/cm, respectively. The initial ion concentrations for the soma model were set to be consistent with current-clamp recording conditions (i.e., 140 mM [Na+]o, 4.5 mM [Na+]i, 5 mM [K+]o, and 130 mM [K+]i), so was the resting membrane potential Vm of −64.3 mV (Shinoda et al. 2010). The initial ion concentrations for the afferent ending model (145 mM [Na+]o, 4 mM [Na+]i, 6.3 mM [K+]o, and 155mM [K+]i) were determined when the model reached equilibrium condition at Vm of −65 mV, a potential well within the range of Vm values recorded from distal axons of small-diameter DRG neurons (−68 to −53 mV) (Vasylyev and Waxman 2012). Simulations were run at 30°C to approximate the experimental conditions of in vitro single-fiber recordings. Rate constants of voltage-dependent channels and Na+-K+-ATPase were multiplied by a temperature factor [i.e., Q10(t − 24)/10]. The Q10 values listed in Table 2 were adapted from Schild et al. (1994). The numerical error tolerance in NEURON was set at 10−5.

Table 2.

Temperature factor Q10 for voltage-gated channels and Na+-K+-ATPase

Rate Parameters Q10
NaV1.6 forward and reverse 1.5
NaV1.7 forward 2.3
NaV1.7 reverse 1.5
NaV1.8 and 1.9 m 2.3
NaV1.8 and 1.9 h 1.5
KA p, q 1.93
KD x, y 1.93
KS n 1.4
Na+-K+-ATPase Imax 1.14

Imax, maximum current; m and h, Hodgkin-Huxley typed voltage-gating parameters for sodium channels; p and q, Hodgkin-Huxley typed voltage-gating parameters for KA current; x and y, Hodgkin-Huxley typed voltage-gating parameters for KD current; n, Hodgkin-Huxley typed voltage-gating parameters for KS current.

Data Recording and Analysis

APs were recorded extracellularly using a low-noise AC differential amplifier. Activity was monitored on-line, filtered (0.3 to 10 kHz), amplified (×10,000), digitized at 20 kHz using a 1401 interface (CED, Cambridge, UK), and stored on a PC. APs were discriminated off-line using Spike 2 software (CED). To avoid erroneous discrimination, no more than two clearly discriminable units in any record were studied. The stretch response threshold was defined as the force that evoked the first AP during ramped stretch. SRFs are presented as bins of evoked APs (0–57, 57–113, 113–170 mN) by the ramped colorectal stretch. SRFs were normalized to the respective maximum binned spike number in control (baseline) tests. To facilitate comparisons, responses of stretch-sensitive afferents are also presented as total numbers of APs during ramped stretch. Data are presented throughout as means ± SE. One-way and two-way ANOVAs or repeated measures were performed as appropriate using SigmaPlot version 11.0 (Systat Software, San Jose, CA). Bonferroni post hoc multiple comparisons were performed when F values for main effects were significant. Differences were considered significant when P < 0.05 (denoted by *).

RESULTS

The Presence of NaV1.6 in Colorectal Afferents

Portions of retrogradely labeled colorectal DRG neurons showed positive immunostaining for NaV1.6 and NaV1.7 (Fig. 2A). Colocalization of YFP (expressed in sensory neurons driven by SNS-Cre) with NaV1.6- and NaV1.7-immunoreactivity in the colorectum suggests the presence of NaV1.6 and NaV1.7 at distal colorectal afferent endings (Fig. 2B). The staining pattern of NaV1.6 is not homogenous along YFP-positive nerve fibers, but rather clusters at focal regions along the axons (see a three-dimensional reconstruction in Supplemental Video S1; the online version of this article contains supplemental data), presumably regions of spike initiation, which is consistent with the sporadic NaV1.6 immunoreactivity in other sensory nerve terminals (Hossain et al. 2005). Both circular and longitudinal smooth muscle layers showed low-intensity staining of NaV1.6, suggesting the presence of some NaV1.6 protein in smooth muscle fibers. In some sections of colorectal tissue, we noted positive NaV1.6 staining that was YFP-negative, likely contributed to by NaV1.6 in efferent endings or enteric neurons (data not shown). Western blots confirmed the presence of NaV1.6 protein in both the PN and L6 DRG, along with NaV1.7 (Fig. 2C). However, the two channels were barely detectable in protein extracts from colon because of relatively low content of neuronal proteins (stained by PGP9.5) in the tissue homogenates.

Effect of TTX on Stretch-Sensitive Colorectal Afferents

As illustrated in Fig. 3A, TTX was applied to the mucosal surface at the receptive field (gray square surrounding “1” on the colorectal mucosal surface), and afferent responses to ramped stretch (0 to 170 mN @ 5 mmHg/s) were assessed prior to and after TTX application and again after washout. TTX at concentrations of 1 μM and 3 μM did not affect responses to stretch (Fig. 3B, F2,8 = 0.12, P = 0.89; Fig. 3C, F2,10 = 2.2, P = 0.16). In contrast, 10 μM TTX inhibited virtually all afferent responses to stretch (Fig. 3D, F2,24 = 28.3, P < 0.001; post hoc comparison, P < 0.001 for TTX vs. control). In additional experiments (n = 7), a bile salt solution (0.25%) was applied locally to the receptive field for 5 min (to increase mucosal permeability) followed by 1 μM TTX application. Application of bile salts significantly increased the response of afferents to stretch (Fig. 3E, F3,36 = 14.9, P < 0.001; post hoc comparison vs. control, P = 0.023), and subsequent application of 1 μM TTX inhibited the response (post hoc comparison vs. control, P = 0.028). The corresponding response thresholds to stretch were unaffected by 1 μM or 3 μM TTX (Fig. 3B, inset, F2,8 = 0.17, P = 0.84; Fig. 3C, inset, F2,10 = 0.84, P = 0.46), but response threshold was significantly increased by 10 μM TTX (Fig. 3D, inset, F2,24 = 24.9, P < 0.001) and 1 μM TTX following bile salts (Fig. 3E, inset, F3,18 = 10.6, P < 0.001; post hoc comparison, P = 0.003 for TTX vs. control). TTX was effectively removed after 15 min of wash (P > 0.05, control vs. wash, for all comparisons).

Fig. 3.

Fig. 3.

Localized mucosal application of tetrodotoxin (TTX) on a stretch-sensitive colorectal afferent ending. A: representative responses to ramped circumferential stretch (0–170 mN @ 5 mN/s) of two stretch-sensitive colorectal afferents (ctrl) and their respective receptive fields (RF 1 and 2) on the flattened colorectal surface. RF 1 was isolated (illustrated by the gray square surrounding RF 1) and exposed to TTX (10 μM); the response was retested after 15 min wash (wash). BD: TTX at 1 and 3 μM modestly, but not significantly, reduced afferent responses to stretch, as shown in B and C, respectively, whereas 10 μM TTX significantly attenuated most, if not all, afferent responses to stretch (D). E: application of 0.25% bile salts significantly increased afferent responses to stretch, and subsequent application of 1 μM TTX significantly attenuated afferent responses. TTX did not increase the response threshold to stretch at 1 or 3 μM (insets in B and C, in mN), but thresholds are significantly increased after 10 μM TTX (inset in D), as well as 1 μM TTX following bile salts application (inset in E). F: responses to stretch were quantified as total number of action potentials during ramped stretch, normalized to control (=1) and are summarized. *P < 0.05.

Responses to stretch at different concentrations of TTX are summarized in Fig. 3F (total spike numbers were normalized to control). When applied to the mucosal surface, the effect of TTX appears to be concentration dependent, but was significant only at the greatest concentration tested (Fig. 3F, 10 μM; F4,47 = 15.6, P < 0.001, post hoc comparison, 10 μM TTX vs. all others, P < 0.02). Application of bile salts (0.25%) significantly increased the responses of afferents to stretch (Fig. 3F, F3,18 = 17.4, post hoc comparison, bile salts vs. control, P = 0.009), and subsequent TTX application at a lower concentration (1 μM) inhibited the response (post hoc comparison, 1 μM TTX vs. control, P = 0.006). There was no significant difference between control and washout (post hoc comparison, P > 0.3).

Effect of NaV1.6 and NaV1.7 Antagonists on Stretch-Sensitive Colorectal Afferents

Both NaV1.6 and NaV1.7 channels are present in primary afferents and are blocked effectively by TTX with comparable EC50 values (Table 3). Thus we assessed the effects of NaV1.6 and NaV1.7 subtype-selective blockers on afferent responses to stretch.

Table 3.

EC50 values of subtype-selective blockers of voltage-gated sodium channels

EC50 NaV1.6 NaV1.7 NaV1.8 Reference
TTX 1–6 4 60,000,000 Catterall et al. 2005
PTX 26 0.3 146 Schmalhofer et al. 2008
mCtxG 680 >100,000 >100,000 Wilson et al. 2011
mCtxP 100 >100,000 >100,000 Wilson et al. 2011
A803 6,740 6,740 140 Jarvis et al. 2007

Values are in nM. TTX, tetrodotoxin; PTX, ProTx-II; mCtxG, μ-conotoxin GIIIa; mCtxP, μ-conotoxin PIIIa; A803, A803467.

As evidenced in Fig. 3, the colon mucosa impedes the diffusion of xenobiotics, including channel-blocking molecules. To avoid concerns related to diffusion across the mucosa and interpretation of results, we conducted the following pharmacological studies with the serosal side of the colorectum facing up. When applied to the serosal surface, 1 μM TTX effectively abolished afferent responses to stretch (Fig. 4A, F2,36 = 30.7, P < 0.001, post hoc comparison vs. control, P < 0.001) and increased the response threshold (inset, F2,18 = 9.2, P = 0.002, post hoc comparison vs. control, P = 0.003). The selective NaV1.7 antagonist PTX at the same concentration (1 μM) did not affect either afferent responses (Fig. 4B, F2,32 = 0.07, P = 0.93) or response threshold (inset, F2,16 = 0.08, P = 0.93). However, PTX at 3 μM, a concentration greater than the EC50 for blocking NaV1.6, effectively attenuated afferent responses (Fig. 4C, F2,20 = 5.9, P = 0.02, post hoc comparison vs. control, P = 0.04) and increased response threshold (inset, F2,10 = 9.6, P = 0.005, post hoc comparison vs. control, P = 0.009). Although blockage of NaV1.7 by PTX is almost irreversible (Johnson et al. 2007), the response to stretch recovered to control completely after only 15 min wash-out (post hoc comparison vs. control, P = 0.95), a time at which NaV1.7 should remain in a blocked state. This strongly suggests that the attenuation of afferent responses to stretch by 3 μM PTX arises through blocking NaV1.6, not NaV1.7. In addition, the selective NaV1.6 antagonist mCtxG (17 μM) significantly reduced afferent responses to stretch (Fig. 4D, F2,36 = 15.2, P < 0.001, post hoc comparison vs. control, P < 0.001), and increased the response threshold (inset, F2,18 = 6.9, P = 0.006, post hoc comparison vs. control, P = 0.008). Data are summarized (total spike number) in Fig. 4E. TTX at 1 μM significantly reduced the response to stretch, whereas the selective NaV1.7 blocker PTX at the same concentration (1 μM) did not (Fig. 4E, F5,99 = 23.1, P < 0.001; post hoc comparison, TTX vs. control, P < 0.001, 1 μM PTX vs. control, P > 0.9). However, 3 μM PTX effectively reduced the responses to stretch (post hoc comparison, 3 μM PTX vs. control, P < 0.001), likely from its nonspecific blocking effect on NaV1.6. In addition, the selective NaV1.6 blocker mCtxG (17 μM) significantly reduced the afferent responses to stretch (post hoc comparison, mCtxG vs. control, P < 0.001), suggesting an essential role of NaV1.6 in colorectal afferent encoding to stretch.

Fig. 4.

Fig. 4.

Localized serosal application of TTX and selective NaV1.6 and NaV1.7 blockers on stretch-sensitive colorectal afferent endings. A: when applied from the serosal side, 1 μM TTX significantly attenuated most, if not all, afferent responses to stretch and increased response threshold (inset). B: in contrast, the NaV1.7 blocker ProTX II (PTX; 1 μM) did not affect afferent responses to stretch or response threshold (inset). C: PTX at 3 μM, a concentration in excess of its EC50 for NaV1.6, significantly attenuated the response to stretch and increased response threshold (inset). D: in contrast, the selective NaV1.6 blocker μ-conotoxin GIIIa (mCtxG; 17 μM) significantly reduced the response to stretch and increased response threshold (inset). E: normalized responses to stretch (total spike number) were significantly reduced by TTX, mCtxG and 3 μM PTX, but not by 1 μM PTX. *P < 0.05.

To verify the finding, the aforementioned selective NaV1.6 and NaV1.7 antagonists were applied to the mucosal surface of the colorectum (3 μM PTX and 50 μM mCtxG), as was another NaV1.6 antagonist from the μ-conotoxin family, mCtxP (10 μM). When applied from the mucosal side, 3 μM PTX did not affect either afferent responses (Fig. 5A, F2,14 = 2.76, P = 0.1) or response threshold (inset, F2,14 = 1.96, P = 0.18) to stretch. In contrast, both mCtxG (50 μM) and mCtxP (10 μM) significantly reduced afferent responses to stretch (Fig. 5B, F2,12 = 28.1, P < 0.001, post hoc comparison vs. control, P < 0.001; Fig. 5C, F2,14 = 19.4, P < 0.001, post hoc comparison vs. control, P < 0.001) and increased the response threshold (Fig. 5B, inset, F2,12 = 9.6, P = 0.003; Fig. 5C, inset, F2,14 = 5.9, P = 0.014). The summarized data in Fig. 5D are consistent with the findings presented in Fig. 4 in which antagonists were applied from the serosal side: both NaV1.6 blockers (mCtxG and mCtxP) significantly reduced the response to stretch, whereas the NaV1.7 blocker PTX did not (F4,50 = 13.3, P < 0.001, post hoc comparison, mCtxG vs. control, P < 0.001, mCtxG vs. control, P < 0.001, PTX vs. control, P = 0.09).

Fig. 5.

Fig. 5.

Localized mucosal application of selective NaV1.6 and NaV1.7 blockers on stretch-sensitive colorectal afferent endings. A and B: when applied from the mucosal side, the NaV1.7 blocker PTX (3 μM) did not affect afferent responses to stretch (A) or response threshold (A, inset), whereas the selective NaV1.6 blocker mCtxG (50 μM) significantly reduced responses to stretch (B) and increased response threshold (B, inset). C: another selective NaV1.6 blocker, μ-conotoxin PIIIa (mCtxP; 10 μM), similarly reduced responses to stretch and increased response threshold (inset). D: normalized responses to stretch (total spike number) were significantly reduced by mCtxG and mCtxP, but not by PTX. The response to stretch after blocking NaV1.6 with mCtxG was significantly lower than after blocking NaV1.7 by PTX. *P < 0.05.

Effect of NaV1.8 Antagonist A803 on Stretch-Sensitive Colorectal Afferents

When applied from the serosal side, the selective NaV1.8 antagonist A803 did not reduce afferent responses to stretch at either 1 μM (Fig. 6A, F2,20 = 1.37, P = 0.3) or 10 μM concentrations (Fig. 6C, F2,24 = 2.97, P = 0.09). Similarly, 3 μM A803 applied from the mucosal side did not affect responses to stretch (Fig. 6B, F2,10 = 4.1, P > 0.05), whereas a greater concentration (10 μM) slightly, but significantly, attenuated responses (Fig. 6D, F2,20 = 6.0, P = 0.009, post hoc comparison vs. control, P = 0.01). Response threshold was unaffected by A803, whether applied from the serosal (Fig. 6A, inset, F2,10 = 0.38, P = 0.69; Fig. 6C, inset, F2,12 = 1.56, P = 0.25) or mucosal side (Fig. 6B, inset, F2,10 = 0.39, P > 0.5; Fig. 6D, inset, F2,20 = 0.74, P > 0.4). Data from serosal and mucosal applications of A803 are summarized (total spike number) in Fig. 6, E and F, respectively; there were no significant differences between control, A803 and wash (Fig. 6E, F3,35 = 2.67, P = 0.06; Fig. 6F, F3,35 = 1.85, P = 0.16). Interestingly, 1 μM A803 applied from the serosal side, a concentration well below the EC50 values for blocking NaV1.6 or NaV1.7 (Table 3), tended to increase the afferent responses to stretch.

Neuron Membrane Models for Colorectal Afferent Endings and c-DRG Somata

Details of the model structures and the ms channel are illustrated in Fig. 1 and described in methods. In addition, the simulations also include four voltage-gated Na+ conductances, three voltage-gated K+ conductances and the Na+-K+-ATPase pump; their maximum conductances and the pump current are presented in Table 4. The sodium channel conductance in the c-DRG soma model was estimated from the peak sodium current (∼5 nA) recorded from c-DRG neurons (Beyak et al. 2004). The density of NaV1.7 and NaV1.8 channels at the siz in the afferent ending model were assigned to be 50 times the density at the soma, an estimation consistent with recent studies indicating 19–60 times sodium channel density at the AIS than at the soma (Baranauskas et al. 2013; Hu et al. 2009; Kole et al. 2008).

Table 4.

Maximum ion channel conductance or pump current

Afferent Ending
Gmax or Imax trsd Soma Soma with NaV1.6 mid siz pas
ms, pA/μm2 1.8
NaV1.6, pS/μm2 1,400 2,800 56
NaV1.7, pS/μm2 2,000 4,000 80 80
NaV1.8 pS/μm2 5,000 10,000 200 200
NaV1.9, pS/μm2 2 4 1 1
KA, pS/μm2 450 900 18 18
KD, pS/μm2 400 800 16 16
KS, pS/μm2 40 210 380 40 20 20
Na+-K+-ATPase, pS/μm2 2.5 2.5 2.5 2.5 0.05 0.05

Gmax, maximum conductance; ms, mechanosensitive channel; trsd, transducer zone; mid, middle section; siz, spike-initiation zone; pas, passive compartment.

Plotted in Fig. 7A are NaV1.6, 1.7, 1.8 and 1.9 currents calculated from voltage clamp simulations (−80 to 70 mV). Different from other channels in the model that use Hodgkin-Huxley equations, NaV1.6 and NaV1.7 were simulated by Markov state models with corresponding diagrams plotted in Fig. 7B. The activation and inactivation channel conductances as functions of membrane voltage (Fig. 7C) were derived by simulating the corresponding single-electrode voltage clamp protocols in whole-cell configurations. Compared with NaV1.7, the activation curve of NaV1.6 is shifted in a hyperpolarized direction, and the inactivation curve in a depolarized direction, suggesting greater open probability during AP generation.

Fig. 7.

Fig. 7.

Simulation of voltage-gated sodium channel subtype (NaV1.6 to 1.9) in DRG somata. A: Na+ currents (equal maximum membrane conductance of 10 pS/μm2) evoked by a single-electrode voltage-clamp protocol (−100-70 mV). B: Markov-type state models for NaV1.6 and NaV1.7 are illustrated to recapitulate their respective gating features [i.e., persistent and resurgent current (dark arrows) and complete inactivation (gray arrow in A)]. C: the activation and inactivation functions of membrane voltage are from normalized peak conductance following simulations of activation and inactivation voltage-clamp protocols.

NaV1.6 Is Critical for Repetitive Spiking

The simulated encoding response of a colorectal afferent ending to stepped circumferential stretch is presented in Fig. 8A. A stepped stretch of 50 mN induced tonic AP generation at the siz, which closely correlated with the transient increase in [Na+]i and transient decrease in [K+]i. Because experimental studies suggested the absence of NaV1.6 currents in small-diameter DRG neurons (Cummins et al. 2005), the NaV1.6 conductance in the c-DRG soma model was set to zero (Table 4). In contrast to the tonic firing in the afferent ending simulation (Fig. 8A), the c-DRG soma model under a stepped current clamp simulation (500 ms) did not fire repetitively, even at a stimulation current three times rheobase (Fig. 8B). However, after adding a NaV1.6 conductance (1/50th of NaV1.6 conductance at siz) in the c-DRG soma model, the soma was able to fire tonically at a stepped current stimulation slightly greater than rheobase (×1.1) as shown in Fig. 8C. In addition, after adding the NaV1.6 conductance to the soma model, the rheobase decreased from 134.5 pA in Fig. 8B to 22.9 pA in Fig. 8C. In contrast, adding the same amount of NaV1.7 or NaV1.8 conductance to the soma model did not change the firing pattern (data not shown, but almost identical to Fig. 8B) other than the slight decrease of rheobase to 116.3 pA and 129.5 pA, respectively.

Fig. 8.

Fig. 8.

Modeled action potential firing patterns in the afferent ending (A) and DRG somata (B). A: the afferent ending model was stimulated by a stepped stretch protocol (50 mN for 15-s duration), which evoked repetitive action potential generation at the siz. The simulation recapitulates the profound changes of intracellular Na+ ([Na+]i) and K+ concentrations ([K+]i) at the siz that correlate with changes in membrane potential (V). B: in contrast, the DRG soma model, when stimulated by a stepped inward current, did not fire repetitively. C: when a NaV1.6 conductance was inserted into the model, the soma model started to fire repetitively at stimulus intensities above rheobase.

Simulation of Afferent Responses to Ramped Stretch after NaV Channel Blockade

During the slow ramp protocol (0 to 170 mN @ 5 mN/s), circumferential stretch deformed the colorectum homogeneously, which was recorded for post hoc measurement of the circumferential λ as displayed in Fig. 9A. To mimic the ramped stretch protocol in single-fiber recordings (Figs. 36), the ramped force (0 to 170 mN in 34 s) and λ were used to drive AP generation at the siz in the afferent ending model with results displayed in Fig. 9B. To save computational time, the falling phase of the ramp, which mirrored the rising phase (34 s) in the experimental protocol, was shortened in duration (6 s) in the simulation. Pharmacological blockage of channels by TTX was mimicked by gradual reduction of maximum conductances of both NaV1.6 and NaV1.7 in the model (by 15%, 30% and 50%). In the simulation, the response to stretch is progressively reduced and completely inhibited by increasing blockage of both NaV1.6 and NaV1.7 conductances, simulating the effect of TTX (Fig. 9B). The pharmacological blockage of subtype-selective NaV channels was simulated by reducing the corresponding maximal conductance by 50% (Fig. 9C). Computational simulations of the total number of spikes evoked by the ramped stretch stimulus are summarized in Fig. 9D, which favorably agrees with the findings from single-fiber studies (i.e., NaV1.6 is necessary for the encoding of colorectal afferent endings to stretch, whereas blockage of NaV1.7 does not remarkably alter the firing pattern). Consistent with the experimental findings using 1 μM A803 (Fig. 6E), blocking NaV1.8 in the model did not reduce afferent responses to stretch, but instead slightly increased firing.

Fig. 9.

Fig. 9.

Simulation of afferent ending response to ramped colorectal stretch. A: colorectal deformation during slow ramped stretch (0 to 170 mN @ 5 mN/s) was measured experimentally, quantified as a circumferential λ, and is plotted. B and C: the experimentally measured mean stretch was used to drive the afferent ending model for action potential generation [TTX (B) and NaV1.7, 1.6, and 1.8 (C)]. D: total spike numbers are summarized. The effect of TTX was simulated by reducing the maximum conductance of both NaV1.6 and NaV1.7, and effects of subtype-selective NaV blockers were simulated by reducing the corresponding NaV conductance. Selective blockage of NaV1.7 or NaV1.8 conductances did not markedly reduce afferent responses to stretch, whereas blockage of NaV1.6 either selectively or by TTX completely inhibited the response.

DISCUSSION

The present study demonstrates that a TTX-sensitive current underlies the tonic encoding of ramped colorectal stretch by a group of unmyelinated sensory afferent endings innervating mouse colorectum. Prior studies on dissociated “nociceptive” DRG neurons suggested that TTX-resistant currents dominate the inward sodium current that drives AP generation (Blair and Bean 2002) because TTX typically did not inhibit AP initiation in DRG somata (Choi and Waxman 2011). Afferents innervating the cornea have been shown to resist blockage by TTX and cooling (Brock et al. 1998; Carr et al. 2003), but a direct effect of TTX on visceral afferent endings has rarely been studied. Andresen et al. (1994) reported that intraluminal perfusion of 40 nM TTX did not markedly affect the encoding function of A-type aortic baroreceptor endings to intraluminal pressure, which is unexpected because A-type baroreceptor somata generally exhibit a TTX-sensitive current (Li and Schild 2007). We found that mucosal application of TTX (10 μM) reversibly inhibited the response of colorectal afferents to stretch, whereas lower concentrations of TTX (1 μM, 3 μM) did not, suggesting that the 40 nM concentration of TTX used by Andresen et al. was insufficient. An alternative interpretation is that the effect of TTX reported here resulted from spillover/leakage and blockage of the PN. This is unlikely based upon 1) the large volume of bath solution (>200 ml) relative to the volume of the TTX solution (150 μl); and 2) the observation that responses to stretch of afferents in the same record with receptive fields outside the locus of TTX application were unaffected (e.g., fiber 2 in Fig. 3A).

The high TTX concentration required to block responses is likely due to epithelial tight-junctions within the colon mucosa that effectively prevent passive diffusion of large molecules (Camilleri et al. 2012). For example, when applied to the colonic mucosa, 300 μM cyclic GMP was required to attenuate colorectal afferent responses to stretch (Feng et al. 2013). However, after pretreating the colorectal mucosa with 0.25% bile salt solution, a detergent that increases gut permeability (Stenman et al. 2013), TTX completely inhibited afferent responses to stretch at 1 μM, a concentration routinely used in patch-clamp studies on tissue slices to block TTX-sensitive currents (Gassner et al. 2009). In addition to increasing gut permeability, bile salts also have a direct effect on sensory afferents (Lieu et al. 2014), which could account for the significant increase in afferent responses to stretch after bile salts application. On the other hand, chemicals applied from the serosal surface appear to bypass the mucosal barrier without affecting the baseline response of the afferents. We documented that chemicals applied to the serosal surface can reach nerve endings at concentrations comparable to those used in patch-clamp studies (e.g., 1 μM TTX effectively blocked afferent responses to stretch).

At the outset, we confirmed the presence of both NaV1.6 and NaV1.7 in colorectal afferent endings and examined the roles of NaV1.6, NaV1.7 and NaV1.8 using subtype-selective antagonists. We excluded study of NaV1.1 (no clear role in nociceptors), NaV1.3 (absent in adult DRG neurons), and NaV1.9 (lack of selective antagonists). We chose PTX as the NaV1.7 antagonist due to its low EC50 (0.3 nM) compared with its EC50s for NaV1.6 (26 nM) and NaV1.8 (146 nM) (Schmalhofer et al. 2008). The two μ-conotoxin NaV1.6 antagonists, mCtxG and mCtxP, have EC50 values for NaV1.6 (0.68 nM for mCtxG and 0.1 nM for mCtxP) orders of magnitude less than for NaV1.7 or NaV1.8 (>100 nM) (Wilson et al. 2011).

NaV1.7, which has been documented as important in pain sensation, was unexpectedly found not necessary for encoding tonic spiking of PN afferent fibers to noxious colorectal stretch. Serosal application of 1 μM PTX did not affect responses to stretch, whereas the same concentration of TTX effectively blocked responses, even though TTX has a much higher EC50 to block NaV1.7 than PTX (Table 3). A greater concentration of PTX (3 μM) slightly, but significantly, attenuated responses to stretch, likely due to its nanomolar EC50 for NaV1.6. A prior study showed that PTX blocks both NaV1.6 and NaV1.7 at nanomolar concentrations, but the blockage of NaV1.6 by PTX has a significantly higher off-rate than blockage of NaV1.7, which accounts for the reversible blockage of NaV1.6 by PTX and almost irreversible blockage of NaV1.7 (Johnson et al. 2007). The fact that the effect of 3 μM PTX on the response to stretch was reversible (response recovered completely after only 15 min wash-out, when NaV1.7 should have remained blocked) strongly suggests the involvement of NaV1.6 and not NaV1.7 in tonic spiking. Further, mCtxG at a concentration that selectively blocks NaV1.6 significantly reduced afferent responses to stretch, indicating a necessary role of NaV1.6 in encoding tonic spiking by stretch-sensitive colorectal afferent endings. An alternative interpretation of the blocking effect by NaV1.6 antagonist is the possible blockage of AP propagation down the axon when applied from the serosal side. But blockage of propagation by our pharmacological scheme seems unlikely, because serosal application of selective NaV1.7 antagonist PTX did not seem to block the afferent response to stretch, whereas a prior study clearly indicated that NaV1.7, not NaV1.6 is critical to AP propagation in unmyelinated C fibers (Schmalhofer et al. 2008). We repeated the experiments with PTX and mCtxG applied from the mucosal side at greater concentrations to overcome the mucosal barrier and also tested another selective NaV1.6 blocker, mCtxP; results confirmed that NaV1.6, not NaV1.7, is necessary for tonic spiking by colorectal afferent endings.

The important role of NaV1.6 in tonic spiking is also supported by recordings from DRG somata. For example, small interfering RNA knockdown of NaV1.6 effectively reduced the proportion of DRG somata that fire repetitively (Xie et al. 2013). NaV1.6 currents are restricted to medium-to-large-diameter DRGs (Cummins et al. 2005), whereas small-diameter DRG somata lack NaV1.6 currents and usually do not fire tonically during stepped current stimulation (Hillsley et al. 2006; Huang et al. 2013; Shinoda et al. 2010). In addition, complementary modeling of a colorectal afferent ending and a soma recapitulated the experimental findings reported here and by others, collectively supporting a necessary role for NaV1.6 in tonic spiking at both afferent endings and somata. Interestingly, both Xie et al. (2013) and we noted positive NaV1.6 immunostaining in some small-diameter DRG somata, which are unlikely to contribute to the membrane current (Cummins et al. 2005).

Perhaps the different spiking properties arise from greater sodium channel densities in afferent endings (50 times in our simulation) than in their somata. Nonetheless, the modeled soma responded tonically at a stimulus intensity slightly greater than rheobase after adding a small NaV1.6 conductance (1/50th the conductance in siz). Thus the lack of tonic firing in c-DRG somata is likely due to the absence of a NaV1.6 conductance, not to the relative low density of other types of sodium channels (e.g., NaV1.7 and NaV1.8) in somata. Different channel compositions at different regions of a same neuron have been documented in AIS in the central nervous system; NaV1.6 channels are clustered at the distal AIS and absent in the soma membrane, whereas NaV1.2 is clustered at the proximal AIS (Baranauskas et al. 2013; Bender and Trussell 2012). The present findings suggest a mechanism/channel to explain differences in neural spiking characteristics between DRG somata and their sensory endings in end organs. Hence, excitability data recorded from dissociated DRG neurons likely do not reflect actual encoding at their sensory endings and thus need to be interpreted with caution.

The selective NaV1.8 antagonist A803 was also studied, but did not significantly affect afferent responses to stretch when applied either to the serosal or mucosal colorectal surface. Closer examination of the data revealed that colorectal afferents were differentially affected by A803; responses to stretch in some afferents were attenuated, whereas in others enhanced. In particular, serosal application of 1 μM A803, a concentration that selectively blocks NaV1.8, tended to increase afferent responses to stretch (P = 0.06). Interestingly, model simulation also showed a slight increase in firing after blocking NaV1.8. In fact, the role of NaV1.8 in nociception and pain remains unclear (Knapp et al. 2012). Given the variability noted, we focused on TTX-sensitive channels NaV1.6 and NaV1.7, leaving NaV1.8 for subsequent study. Similarly, we included in our simulation model the gating formulas for an NaV1.9 conductance, which appears to a play significant role during repeated noxious colon distension (Hockley et al. 2014) and is beyond the focus of the current study.

In previous sensory neuron models, NaV1.6 and NaV1.7 conductances were represented together as one TTX-sensitive conductance by a Hodgkin-Huxley styled formulation (Amir and Devor 2003; Baker 2005; Kovalsky et al. 2009; Schild et al. 1994; Tigerholm et al. 2014), which cannot simulate the unique gating features of NaV1.7 [i.e., complete inactivation and slow recovery from inactivation (Gurkiewicz et al. 2011)]. The computational models employed here incorporated individual Markov models for NaV1.6 (Khaliq et al. 2003) and NaV1.7 (Gurkiewicz et al. 2011), previously verified with experimental data to capture their contrasting gating features (i.e., rapid vs. slow repriming, presence vs. absence of sustained current, and difference in inactivation voltage). The simulation revealed that the contrasting roles for NaV1.6 and NaV1.7 in tonic firing are likely caused by those aforementioned differences in channel gating properties.

Our modeling also incorporated a novel ms channel that drives AP generation, which has not been reported previously. The rate constants between channel gating states were formulated as exponential functions of membrane tension, resulting in a Boltzmann-like steady-state channel open probability, consistent with prior experimental and theoretical studies (Hao and Delmas 2010; Haselwandter and Phillips 2013; Wiggins and Phillips 2005). Membrane tension at the afferent ending is linked to bulk circumferential colorectal stretch by a lumped parametric model that is routinely used to simulate passive mechanical properties of biological tissues (Feng and Gan 2004). The ms current evoked by a stepped stretch was consistent with a slowly-adapting ms current recorded in DRG neurons (Rugiero et al. 2010). The formulation also allows simulation of other types of ms (e.g., rapidly adapting) currents in future investigations when the scope of study extends beyond stretch-sensitive afferents.

The current computational model borrowed heavily from prior computational studies, including adoption of parameters for the voltage-gated channels and Na-K pump. The maximal conductances of sodium channels at the siz was set to be 50 times the conductance at the soma; the conductance at the soma was determined from voltage-clamp data recorded on colorectal DRG neurons (Beyak et al. 2004). Parameters that were adjusted include maximal conductances of potassium channels necessary for the repolarization of the membrane depolarization, and gating parameters associated with ms channels. We did not systematically adjust model parameters or conduct extensive sensitivity studies here, but instead focused on uncovering the differential roles of NaV1.6, NaV1.7 and NaV1.8 conductances in afferent encoding of ramped stretch, whose gating formulas were rigorously verified in previous studies (see Table 1 for references). Also, the geometry of the afferent ending model is intended to simulate the electrotonic character of free nerve endings, a morphological feature of the majority of the colorectal afferent terminals. However, some colorectal afferent endings appear to have lamina- and array-like morphologies, which will not be adequately simulated by the current model geometry and need to be addressed in future studies.

In summary, we provide experimental and computational evidence for a necessary role of a NaV1.6 current in neural encoding of stretch-sensitive colorectal afferents. Immunohistochemistry and Western blotting revealed the presence of NaV1.6 and NaV1.7 at colorectal neuronal endings. Both TTX and selective NaV1.6 antagonists significantly attenuated afferent responses to stretch, whereas a selective NaV1.7 antagonist only slightly reduced the response. Computational Markov type modeling recapitulated the pharmacological findings. A selective NaV1.8 antagonist did not significantly attenuate the responses to stretch, and the exact role of NaV1.8 requires further study. These computational models provide a solid theoretical foundation for future studies of underlying mechanisms of neural encoding in different classes of colorectal afferents in both physiological and pathophysiological conditions.

GRANTS

This study was supported by National Institute of Diabetes and Digestive and Kidney Diseases Grant R01 DK-093525 (G. F. Gebhart).

DISCLOSURES

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

AUTHOR CONTRIBUTIONS

Author contributions: B.F. and G.F.G. conception and design of research; B.F., Y.Z., J.-H.L., and Z.P.W. performed experiments; B.F., Y.Z., J.-H.L., Z.P.W., and G.F.G. analyzed data; B.F., Z.P.W., and G.F.G. interpreted results of experiments; B.F., Y.Z., J.-H.L., and Z.P.W. prepared figures; B.F., Y.Z., and J.-H.L. drafted manuscript; B.F., Z.P.W., and G.F.G. edited and revised manuscript; B.F., Y.Z., J.-H.L., Z.P.W., and G.F.G. approved final version of manuscript.

Supplementary Material

Video S1
Download video file (7.3MB, mp4)

ACKNOWLEDGMENTS

We thank Michael Burcham for assistance in preparation of figures and Dr. Brian Davis for generous gift of transgenic mice that express YFP in sensory afferents (driven by SNS-Cre).

REFERENCES

  1. Amir R, Devor M. Extra spike formation in sensory neurons and the disruption of afferent spike patterning. Biophys J 84: 2700–2708, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Andresen MC, Brodwick M, Yang M. Contrasting actions of cocaine, local anaesthetic and tetrodotoxin on discharge properties of rat aortic baroreceptors. J Physiol 477: 309–319, 1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  3. 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. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Baranauskas G, David Y, Fleidervish IA. Spatial mismatch between the Na+ flux and spike initiation in axon initial segment. Proc Natl Acad Sci U S A 110: 4051–4056, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  5. Bender KJ, Trussell LO. The physiology of the axon initial segment. Annu Rev Neurosci 35: 249–265, 2012. [DOI] [PubMed] [Google Scholar]
  6. Beyak MJ, Ramji N, Krol KM, Kawaja MD, Vanner SJ. Two TTX-resistant Na+ currents in mouse colonic dorsal root ganglia neurons and their role in colitis-induced hyperexcitability. Am J Physiol Gastrointest Liver Physiol 287: G845–G855, 2004. [DOI] [PubMed] [Google Scholar]
  7. Blair NT, Bean BP. 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, 2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Bondarenko VE, Szigeti GP, Bett GC, Kim SJ, Rasmusson RL. Computer model of action potential of mouse ventricular myocytes. Am J Physiol Heart Circ Physiol 287: H1378–H1403, 2004. [DOI] [PubMed] [Google Scholar]
  9. Brierley SM, Jones RC 3rd, Gebhart GF, Blackshaw LA. Splanchnic and pelvic mechanosensory afferents signal different qualities of colonic stimuli in mice. Gastroenterology 127: 166–178, 2004. [DOI] [PubMed] [Google Scholar]
  10. Brock JA, McLachlan EM, Belmonte C. Tetrodotoxin-resistant impulses in single nociceptor nerve terminals in guinea-pig cornea. J Physiol 512: 211–217, 1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Busby RW, Kessler MM, Bartolini WP, Bryant AP, Hannig G, Higgins CS, Solinga RM, Tobin JV, Wakefield JD, Kurtz CB, Currie MG. Pharmacologic properties, metabolism, and disposition of linaclotide, a novel therapeutic peptide approved for the treatment of irritable bowel syndrome with constipation and chronic idiopathic constipation. J Pharmacol Exp Ther 344: 196–206, 2013. [DOI] [PubMed] [Google Scholar]
  12. Camilleri M, Madsen K, Spiller R, Greenwood-Van Meerveld B, Verne GN. Intestinal barrier function in health and gastrointestinal disease. Neurogastroenterol Motil 24: 503–512, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Carnevale NT, Hines ML. The NEURON book. Cambridge MA: Cambridge University Press, 2005, p. xix, 457. [Google Scholar]
  14. Carr RW, Pianova S, Fernandez J, Fallon JB, Belmonte C, Brock JA. Effects of heating and cooling on nerve terminal impulses recorded from cold-sensitive receptors in the guinea-pig cornea. J Gen Physiol 121: 427–439, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Catterall WA, Goldin AL, Waxman SG. International Union of Pharmacology. XLVII Nomenclature and structure-function relationships of voltage-gated sodium channels. Pharmacol Rev 57: 397–409, 2005. [DOI] [PubMed] [Google Scholar]
  16. Chey WD, Lembo AJ, Lavins BJ, Shiff SJ, Kurtz CB, Currie MG, MacDougall JE, Jia XD, Shao JZ, Fitch DA, Baird MJ, Schneier HA, Johnston JM. Linaclotide for irritable bowel syndrome with constipation: a 26-wk, randomized, double-blind, placebo-controlled trial to evaluate efficacy and safety. Am J Gastroenterol 107: 1702–1712, 2012. [DOI] [PubMed] [Google Scholar]
  17. Choi JS, Waxman SG. Physiological interactions between Na(v)1.7 and Na(v)18 sodium channels: a computer simulation study. J Neurophysiol 106: 3173–3184, 2011. [DOI] [PubMed] [Google Scholar]
  18. Christianson JA, Traub RJ, Davis BM. Differences in spinal distribution and neurochemical phenotype of colonic afferents in mouse and rat. J Comp Neurol 494: 246–259, 2006. [DOI] [PubMed] [Google Scholar]
  19. Cummins TR, Dib-Hajj SD, Herzog RI, Waxman SG. Nav1.6 channels generate resurgent sodium currents in spinal sensory neurons. FEBS Lett 579: 2166–2170, 2005. [DOI] [PubMed] [Google Scholar]
  20. Feng B, Brumovsky PR, Gebhart GF. Differential roles of stretch-sensitive pelvic nerve afferents innervating mouse distal colon and rectum. Am J Physiol Gastrointest Liver Physiol 298: G402–G409, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Feng B, Gan RZ. Lumped parametric model of the human ear for sound transmission. Biomech Model Mechanobiol 3: 33–47, 2004. [DOI] [PubMed] [Google Scholar]
  22. Feng B, Gebhart GF. Characterization of silent afferents in the pelvic and splanchnic innervations of the mouse colorectum. Am J Physiol Gastrointest Liver Physiol 300: G170–G180, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Feng B, Kiyatkin ME, La JH, Ge P, Solinga R, Silos-Santiago I, Gebhart GF. Activation of guanylate cyclase-C attenuates stretch responses and sensitization of mouse colorectal afferents. J Neurosci 33: 9831–9839, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Feng B, La JH, Schwartz ES, Gebhart GF. Irritable bowel syndrome: methods, mechanisms, and pathophysiology. Neural and neuro-immune mechanisms of visceral hypersensitivity in irritable bowel syndrome. Am J Physiol Gastrointest Liver Physiol 302: G1085–G1098, 2012a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. Feng B, La JH, Schwartz ES, Tanaka T, McMurray TP, Gebhart GF. Long-term sensitization of mechanosensitive and -insensitive afferents in mice with persistent colorectal hypersensitivity. Am J Physiol Gastrointest Liver Physiol 302: G676–G683, 2012b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Feng B, La JH, Tanaka T, Schwartz ES, McMurray TP, Gebhart GF. Altered colorectal afferent function associated with TNBS-induced visceral hypersensitivity in mice. Am J Physiol Gastrointest Liver Physiol 303: G817–G824, 2012c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  27. Fleidervish IA, Lasser-Ross N, Gutnick MJ, Ross WN. Na+ imaging reveals little difference in action potential-evoked Na+ influx between axon and soma. Nat Neurosci 13: 852–860, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Gassner M, Ruscheweyh R, Sandkuhler J. Direct excitation of spinal GABAergic interneurons by noradrenaline. Pain 145: 204–210, 2009. [DOI] [PubMed] [Google Scholar]
  29. Gurkiewicz M, Korngreen A, Waxman SG, Lampert A. Kinetic modeling of Nav1.7 provides insight into erythromelalgia-associated F1449V mutation. J Neurophysiol 105: 1546–1557, 2011. [DOI] [PubMed] [Google Scholar]
  30. Hao J, Delmas P. Multiple desensitization mechanisms of mechanotransducer channels shape firing of mechanosensory neurons. J Neurosci 30: 13384–13395, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Haselwandter CA, Phillips R. Connection between oligomeric state and gating characteristics of mechanosensitive ion channels. PLoS Comput Biol 9: e1003055, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Hillsley K, Lin JH, Stanisz A, Grundy D, Aerssens J, Peeters PJ, Moechars D, Coulie B, Stead RH. Dissecting the role of sodium currents in visceral sensory neurons in a model of chronic hyperexcitability using Nav1.8 and Nav1.9 null mice. J Physiol 576: 257–267, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Hockley JR, Boundouki G, Cibert-Goton V, McGuire C, Yip PK, Chan C, Tranter M, Wood JN, Nassar MA, Blackshaw LA, Aziz Q, Michael GJ, Baker MD, Winchester WJ, Knowles CH, Bulmer DC. Multiple roles for NaV1.9 in the activation of visceral afferents by noxious inflammatory, mechanical, and human disease-derived stimuli. Pain 155: 1962–1975, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  34. Hodgkin AL. The local electric changes associated with repetitive action in a non-medullated axon. J Physiol 107: 165–181, 1948. [DOI] [PMC free article] [PubMed] [Google Scholar]
  35. Hossain WA, Antic SD, Yang Y, Rasband MN, Morest DK. Where is the spike generator of the cochlear nerve? Voltage-gated sodium channels in the mouse cochlea. J Neurosci 25: 6857–6868, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Hu W, Tian C, Li T, Yang M, Hou H, Shu Y. Distinct contributions of Na(v)1.6 and Na(v)1.2 in action potential initiation and backpropagation. Nat Neurosci 12: 996–1002, 2009. [DOI] [PubMed] [Google Scholar]
  37. 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. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Jarvis MF, Honore P, Shieh CC, Chapman M, Joshi S, Zhang XF, Kort M, Carroll W, Marron B, Atkinson R, Thomas J, Liu D, Krambis M, Liu Y, McGaraughty S, Chu K, Roeloffs R, Zhong C, Mikusa JP, Hernandez G, Gauvin D, Wade C, Zhu C, Pai M, Scanio M, Shi L, Drizin I, Gregg R, Matulenko M, Hakeem A, Gross M, Johnson M, Marsh K, Wagoner PK, Sullivan JP, Faltynek CR, Krafte DS. A-803467, a potent and selective Nav1.8 sodium channel blocker, attenuates neuropathic and inflammatory pain in the rat. Proc Natl Acad Sci U S A 104: 8520–8525, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  39. Johnson JP, Urrutia A, Depry J, Fox T, Gonzalez JE, Neuberger T, Shah S, Silos-Santiago I, Wynn M, Cohen CJ. Isolation of NaV1.7 current in sensory and sympathetic neurons with a selective spider toxin, ProTx-2. Program No 46624/I10. In: 2007 Neuroscience Meeting Planner. San Diego, CA: Society for Neuroscience, 2007. [Google Scholar]
  40. Khaliq ZM, Gouwens NW, Raman IM. The contribution of resurgent sodium current to high-frequency firing in Purkinje neurons: an experimental and modeling study. J Neurosci 23: 4899–4912, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  41. Kiyatkin ME, Feng B, Schwartz ES, Gebhart GF. Combined genetic and pharmacological inhibition of TRPV1 and P2X3 attenuates colorectal hypersensitivity and afferent sensitization. Am J Physiol Gastrointest Liver Physiol 305: G638–G648, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Knapp O, McArthur JR, Adams DJ. Conotoxins targeting neuronal voltage-gated sodium channel subtypes: potential analgesics? Toxins (Basel) 4: 1236–1260, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Kole MH, Ilschner SU, Kampa BM, Williams SR, Ruben PC, Stuart GJ. Action potential generation requires a high sodium channel density in the axon initial segment. Nat Neurosci 11: 178–186, 2008. [DOI] [PubMed] [Google Scholar]
  44. Kovalsky Y, Amir R, Devor M. 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, 2009. [DOI] [PubMed] [Google Scholar]
  45. La JH, Feng B, Schwartz ES, Brumovsky PR, Gebhart GF. Luminal hypertonicity and acidity modulate colorectal afferents and induce persistent visceral hypersensitivity. Am J Physiol Gastrointest Liver Physiol 303: G802–G809, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. La JH, Schwartz ES, Gebhart GF. Differences in the expression of transient receptor potential channel V1, transient receptor potential channel A1 and mechanosensitive two pore-domain K+ channels between the lumbar splanchnic and pelvic nerve innervations of mouse urinary bladder and colon. Neuroscience 186: 179–187, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Li BY, Schild JH. Electrophysiological and pharmacological validation of vagal afferent fiber type of neurons enzymatically isolated from rat nodose ganglia. J Neurosci Methods 164: 75–85, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Lieu T, Jayaweera G, Zhao P, Poole DP, Jensen D, Grace M, McIntyre P, Bron R, Wilson YM, Krappitz M, Haerteis S, Korbmacher C, Steinhoff MS, Nassini R, Materazzi S, Geppetti P, Corvera CU, Bunnett NW. The bile acid receptor TGR5 activates the TRPA1 channel to induce itch in mice. Gastroenterology 147: 1417–1428, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  49. Naliboff BD, Munakata J, Fullerton S, Gracely RH, Kodner A, Harraf F, Mayer EA. Evidence for two distinct perceptual alterations in irritable bowel syndrome. Gut 41: 505–512, 1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Raman IM, Bean BP. Inactivation and recovery of sodium currents in cerebellar Purkinje neurons: evidence for two mechanisms. Biophys J 80: 729–737, 2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Rao S, Lembo AJ, Shiff SJ, Lavins BJ, Currie MG, Jia XD, Shi K, MacDougall JE, Shao JZ, Eng P, Fox SM, Schneier HA, Kurtz CB, Johnston JM. A 12-week, randomized, controlled trial with a 4-week randomized withdrawal period to evaluate the efficacy and safety of linaclotide in irritable bowel syndrome with constipation. Am J Gastroenterol 107: 1714–1725, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. Rugiero F, Drew LJ, Wood JN. Kinetic properties of mechanically activated currents in spinal sensory neurons. J Physiol 588: 301–314, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Rush AM, Dib-Hajj SD, Waxman SG. Electrophysiological properties of two axonal sodium channels, Nav1.2 and Nav1.6, expressed in mouse spinal sensory neurons. J Physiol 564: 803–815, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  54. Schild JH, Clark JW, Hay M, Mendelowitz D, Andresen MC, Kunze DL. A- and C-type rat nodose sensory neurons: model interpretations of dynamic discharge characteristics. J Neurophysiol 71: 2338–2358, 1994. [DOI] [PubMed] [Google Scholar]
  55. Schmalhofer WA, Calhoun J, Burrows R, Bailey T, Kohler MG, Weinglass AB, Kaczorowski GJ, Garcia ML, Koltzenburg M, Priest BT. ProTx-II, a selective inhibitor of NaV1.7 sodium channels, blocks action potential propagation in nociceptors. Mol Pharmacol 74: 1476–1484, 2008. [DOI] [PubMed] [Google Scholar]
  56. Shinoda M, La JH, Bielefeldt K, Gebhart GF. Altered purinergic signaling in colorectal dorsal root ganglion neurons contributes to colorectal hypersensitivity. J Neurophysiol 104: 3113–3123, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Stenman LK, Holma R, Eggert A, Korpela R. A novel mechanism for gut barrier dysfunction by dietary fat: epithelial disruption by hydrophobic bile acids. Am J Physiol Gastrointest Liver Physiol 304: G227–G234, 2013. [DOI] [PubMed] [Google Scholar]
  58. Tigerholm J, Petersson ME, Obreja O, Lampert A, Carr R, Schmelz M, Fransen EA. Modeling activity-dependent changes of axonal spike conduction in primary afferent C-nociceptors. J Neurophysiol 111: 1721–1735, 2014. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Vasylyev DV, Waxman SG. Membrane properties and electrogenesis in the distal axons of small dorsal root ganglion neurons in vitro. J Neurophysiol 108: 729–740, 2012. [DOI] [PubMed] [Google Scholar]
  60. Verne GN, Robinson ME, Vase L, Price DD. Reversal of visceral and cutaneous hyperalgesia by local rectal anesthesia in irritable bowel syndrome (IBS) patients. Pain 105: 223–230, 2003. [DOI] [PubMed] [Google Scholar]
  61. Verne GN, Sen A, Price DD. Intrarectal lidocaine is an effective treatment for abdominal pain associated with diarrhea-predominant irritable bowel syndrome. J Pain 6: 493–496, 2005. [DOI] [PubMed] [Google Scholar]
  62. Wiggins P, Phillips R. Membrane-protein interactions in mechanosensitive channels. Biophys J 88: 880–902, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  63. Wilson MJ, Yoshikami D, Azam L, Gajewiak J, Olivera BM, Bulaj G, Zhang MM. mu-Conotoxins that differentially block sodium channels NaV1.1 through 1.8 identify those responsible for action potentials in sciatic nerve. Proc Natl Acad Sci U S A 108: 10302–10307, 2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Xie W, Strong JA, Ye L, Mao JX, Zhang JM. Knockdown of sodium channel NaV1.6 blocks mechanical pain and abnormal bursting activity of afferent neurons in inflamed sensory ganglia. Pain 154: 1170–1180, 2013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Zhu Y, Lu SG, Gold MS. Persistent inflammation increases GABA-induced depolarization of rat cutaneous dorsal root ganglion neurons in vitro. Neuroscience 220: 330–340, 2012. [DOI] [PMC free article] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Video S1
Download video file (7.3MB, mp4)

Articles from Journal of Neurophysiology are provided here courtesy of American Physiological Society

RESOURCES