Skip to main content
Journal of Neurophysiology logoLink to Journal of Neurophysiology
. 2010 Mar 10;103(5):2532–2543. doi: 10.1152/jn.00506.2009

Dendritic HCN Channels Shape Excitatory Postsynaptic Potentials at the Inner Hair Cell Afferent Synapse in the Mammalian Cochlea

Eunyoung Yi 1, Isabelle Roux 1, Elisabeth Glowatzki 1,
PMCID: PMC2867566  PMID: 20220080

Abstract

Synaptic transmission at the inner hair cell (IHC) afferent synapse, the first synapse in the auditory pathway, is specialized for rapid and reliable signaling. Here we investigated the properties of a hyperpolarization-activated current (Ih), expressed in the afferent dendrite of auditory nerve fibers, and its role in shaping postsynaptic activity. We used whole cell patch-clamp recordings from afferent dendrites directly where they contact the IHC in excised postnatal rat cochlear turns. Excitatory postsynaptic potentials (EPSPs) of variable amplitude (1–35 mV) were found with 10–90% rise times of about 1 ms and time constants of decay of about 5 ms at room temperature. Current–voltage relations recorded in afferent dendrites revealed Ih. The pharmacological profile and reversal potential (−45 mV) indicated that Ih is mediated by hyperpolarization-activated cyclic nucleotide-gated cation (HCN) channels. The HCN channel subunits HCN1, HCN2, and HCN4 were found to be expressed in afferent dendrites using immunolabeling. Raising intracellular cAMP levels sped up the activation kinetics, increased the magnitude of Ih and shifted the half activation voltage (Vhalf) to more positive values (−104 ± 3 to −91 ± 2 mV). Blocking Ih with 50 μM ZD7288 resulted in hyperpolarization of the resting membrane potential (∼4 mV) and slowing the decay of the EPSP by 47%, suggesting that Ih is active at rest and shortens EPSPs, thereby potentially improving rapid and reliable signaling at this first synapse in the auditory pathway.

INTRODUCTION

To perform tasks such as the localization of sound in space, neurons in the auditory pathway are specialized to accurately preserve timing information within sound signals (Oertel 1999; Trussell 1999). Several pre- and postsynaptic mechanisms enabling rapid and reliable transmission at auditory synapses have been described. Presynaptically, large calyceal structures release vesicles from many release sites synchronously, resulting in large excitatory postsynaptic potentials (EPSPs) that reliably activate action potentials (APs) (Schneggenburger and Forsythe 2006). Postsynaptically, rapid kinetics of α-amino-3-hydroxy-5-methyl-4-isoxazolepropionic acid (AMPA) receptors result in brief excitatory postsynaptic currents (EPSCs) (Gardner et al. 1999, 2001; Parks 2000; Raman et al. 1994). Voltage-gated ion channels active near the resting membrane potential decrease the membrane resistance of the postsynaptic neurons and shorten the membrane time constant (Magee and Johnston 2005). This mechanism keeps EPSPs brief and prevents temporal summation of synaptic events. A hyperpolarization-activated cation channel (Ih) has been described in auditory brain stem neurons as one of the ion channels serving this role (Banks and Smith 1992; Golding et al. 1995; Rothman and Manis 2003b). To further shape EPSPs, auditory neurons can receive modulatory inputs that either open receptor-coupled ion channels (Funabiki et al. 1998; Smith et al. 2000) or regulate Ih via G protein coupled signaling pathways (Banks et al. 1993; Yamada et al. 2005).

The first synapse in the auditory pathway, the synapse between the inner hair cell (IHC) and auditory nerve fiber, also uses highly specialized mechanisms to preserve timing information (Fuchs 2005; Glowatzki et al. 2008; Moser et al. 2006). The auditory nerve fiber receives input from only one IHC via a single dendrite and large EPSCs are activated by multivesicular release at this ribbon-type synapse (Glowatzki and Fuchs 2002; Goutman and Glowatzki 2007; Grant et al. 2010; Keen and Hudspeth 2006; Li et al. 2009). Similar to EPSCs recorded from auditory brain stem synapses, AMPA-mediated EPSCs at this synapse are brief (Glowatzki and Fuchs 2002; Grant et al. 2010). However, not much is known regarding the expression pattern of voltage-gated ion channels in afferent dendrites and their involvement in shaping postsynaptic activity. In a first survey, using voltage-clamp recordings from afferent dendrites of the postnatal rat cochlea, we identified several voltage-gated conductances in afferent dendrites (Na+, Ca 2+, K+ conductances) including a hyperpolarization-activated conductance.

Here we focus on the characterization of Ih in IHC afferent dendrites. We find that Ih is mediated by HCN channels. Ih is active at rest and is modulated by intracellular levels of cyclic adenosine monophosphate (cAMP). Ih shortens the EPSP and is thus a good candidate for enabling rapid and reliable signaling at this first synapse in the auditory pathway.

METHODS

Animal protocols were approved by the Johns Hopkins University Animal Care and Use Committee. Rats (Sprague–Dawley; Charles River, Wilmington, MA) were anesthetized (pentobarbital 0.045 mg·g−1, administered intraperitoneally or by isoflurane inhalation) and decapitated and cochleae were quickly removed from temporal bones.

Electrophysiological recordings

Excised apical cochlear turns of 7- to 14-day-old rats were placed into a chamber under an upright microscope (Axioskop2 FS plus, Zeiss, Oberkochen, Germany) and superfused with external solution at 1–3 ml/min (chamber volume ∼2 ml). IHCs and contacting afferent dendrites were visualized on a monitor via a ×40 water immersion objective, ×4 magnification, differential interference contrast optics using a green filter, and a NC 70 Newvicon camera (Dage, MTI, Michigan City, IN). The pipette solution consisted of (in mM): 135 KCl, 3.5 MgCl2, 0.1 CaCl2, 5 EGTA, 5 HEPES, and 0–2.5 Na2ATP; or 135 KCl, 3.5 MgCl2, 0.1 CaCl2, 5 EGTA, 5 HEPES, 4 Na2ATP, and 0.2 Na2GTP; 290 mOsm, pH 7.2 (KOH). In some recordings in which the effect of cAMP was tested, the pipette solution contained (in mM):131 KCl, 1 MgCl2, 5 EGTA, 5 HEPES, 5 Na2ATP, and 10 Na2-phosphocreatine. In addition, a small number of recordings (n = 6) was performed using a pipette solution containing (in mM): 110 K-methanesulfonate, 20 KCl, 5 EGTA, 5 HEPES, 0.1 CaCl2, 5 Na2-phosphocreatine, 4 MgATP, and 0.3 Tris-GTP. No significant differences were found in basic membrane properties such as input resistance and membrane time constant of the afferent fiber between recordings with KCl- or K-methanesulfonate-based pipette solutions. The external solution consisted of (in mM): 5.8 KCl, 155 NaCl, 1.3 CaCl2, 0.9 MgCl2, 0.7 NaH2PO4, 5.6 glucose, and 10 HEPES; 300 mOsm, pH 7.4 (NaOH). Drugs were dissolved daily in the external solution to their final concentrations from frozen stocks. Application of drug solutions was performed using a gravity-driven flow pipette (100 μm diameter) placed near the row of IHCs, connected with a VC-6 channel valve controller (Warner Instruments, Hamden, CT). 4-Ethylphenylamino-1,2-dimethyl-6-methylaminopyrimidinium chloride (ZD7288), 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX), and 4-aminopyridine (4-AP) were purchased from Tocris Bioscience (Ellisville, MO) and tetrodotoxin (TTX) either from Alomone (Jerusalem, Israel) or Sigma (St. Louis, MO). All other chemicals were purchased from Sigma.

Recording pipettes were fabricated from 1 mm borosilicate glass (WPI, Sarasota, FL). Pipettes were pulled with a multistep horizontal puller (Sutter Instrument, San Rafael, CA) and fire-polished (10–15 MΩ). Pipettes were coated with Sylgard (Dow Corning, Midland, MI). Experiments were done at 22–25°C. Recordings were performed with a Multiclamp 700A or 700B amplifier (Molecular Devices, Sunnyvale, CA), pClamp version 9.2, and a Digidata 1322A board, digitized at 50 kHz, and filtered at 10 kHz.

In voltage-clamp mode, series resistance (Rs) was calculated from capacitative current responses to a 10 mV voltage step (−84 to −94 mV). The capacitative current responses were fitted with a sum of two or more exponential equations. The fastest component of the fit was considered to represent the capacitative current for the membrane area near the recording electrode (Cfast) and the slower components for current spreading further along the afferent nerve fiber (Cslow) (Llano et al. 1991). From the fastest component, the voltage-clamp time constant (36 ± 12 μs, n = 38), membrane capacitance Cfast (1.32 ± 0.43 pF), and Rs (30 ± 12 MΩ) were derived. Voltage-clamp data were discarded if Rs was >50 MΩ. Most Ih currents were <300 pA and, with Rs at about 30 MΩ, the estimated voltage error was <9 mV.

Assuming that the dendrite is formed like a cylinder and with a specific capacitance of 1 μF/cm2 and a diameter of 1 μm (most likely an overestimation), the first component corresponds to the dendrite at a length of about 40 μm. This distance covers the extent of the HCN channel expression along the terminal; the HCN specific labeling ceases in the region of the first heminode of the peripheral dendrite, about 50 μm away from the afferent contacts with the IHC. Synaptic currents are generated within <3 μm of the pipette tip as the tip is directly positioned on the bouton ending. Therefore voltage-clamp conditions for recording synaptic currents and HCN channels should be sufficient. The input resistance (Rin) of afferent dendrites (394 ± 253 MΩ, n = 173) was determined in voltage clamp, with voltage steps from −64 to −84 mV.

In current-clamp mode, errors due to Rs were compensated using bridge balance and pipette capacitance neutralization. Membrane voltage responses to −10 pA current steps were fitted with a monoexponential equation and provided a membrane time constant (τm) of 3.97 ± 1.81 ms (n = 10). τm measured in current-clamp mode is larger than τm estimated from voltage-clamp data (0.5 ms) when only Cfast is used to estimate cell capacitance. This happens most likely because in current-clamp mode the effective cell capacitance is not limited to Cfast but also includes Cslow.

Liquid junction potentials (4 mV for KCl-based and 9 mV for K-methanesulfonate-based pipette solution) were corrected off-line. Data were analyzed off-line using pClamp version 9.2 (Axon Instruments, Union City, CA), Minianalysis (Synaptosoft, Decatur, GA), and Origin 7.5 (OriginLab, Northampton, MA). For statistical comparisons Sigmastat 3.5 (Systat Software, San Jose, CA) was used. Statistical significances of irreversible drug effects (ZD7288) were tested using a paired t-test. Effects of ZD7288 on the EPSP waveform (measurements were taken at three different conditions) were tested using one-way repeated measures ANOVA followed by Student–Newman–Keuls test. Effects of reversible drugs (CsCl and BaCl2) were tested using one-way repeated measures ANOVA followed by Student–Newman–Keuls test. Effects of cAMP analogs on Ih amplitude and activation kinetics were tested using two-way repeated measures ANOVA one factor repetition. Effects of cAMP on the Vhalf and slope factor of Ih activation curves were tested using Student's t-test. Values are presented as means ± SD.

Immunolabeling

Cochleae from 9- to 10- and 20- to 21-day-old rats (P9–P10, P20–P21) were perfused through the round and oval windows with cold 4% paraformaldehyde prepared in phosphate buffered saline (PBS), pH 7.4, and then postfixed for 1 h at 4°C under agitation. Additionally, for HCN1 immunodetection, cochleae were rinsed three times in PBS and incubated 15 min in methanol at −20°C. Thereafter, preparations were washed three times in PBS and the cochleae were microdissected to facilitate access of the antibodies to the tissue. Whole-mount preparations were incubated for 1 h at room temperature in a blocking and permeabilizing solution (PBS with 20% of either normal goat serum or donkey serum and 0.3% Triton X-100) and were then incubated overnight at 4°C with the primary antibodies diluted in the same solution. After three 15 min washes in PBS, samples were incubated for 1 h at room temperature with fluorescently labeled secondary antibodies diluted at 1:800 in PBS with 10% of either normal goat serum or donkey serum and 0.15% Triton X-100. Samples were then rinsed once in PBS with 10% of either normal goat serum or donkey serum and 0.15% Triton X-100 and twice with PBS (15 min each, at room temperature) before the organs of Corti were mounted on slides using FluorSave mounting medium (Calbiochem, San Diego, CA). Specific labeling was initially examined with an Axio Observer inverted microscope (Zeiss) and further detailed images were obtained using a confocal laser scanning microscope (LSM 510 META, Zeiss) with ×20 air and ×100 oil objectives (optical section steps of 0.25 and 0.20 microns, respectively). Analysis and reconstruction were carried out using LSM Image Examiner (Zeiss) and Volocity 4.2.1 software (Improvision, Waltham, MA). No labeling was observed when the primary antibodies were omitted. Again, no labeling was observed when the primary antibodies were preabsorbed onto target peptides, except in the stereocilia of the sensory hair cells, which displayed staining using HCN1 and HCN4 antibodies, when the target peptides were used at the recommended concentration or at a 25-fold (HCN1) or 5-fold (HCN4) higher concentration. For HCN2, no test for preabsorption with a peptide was performed. For HCN3 a shorter fixation (15 min) was also tested.

Antibodies

Rabbit polyclonal antibodies against HCN1 and HCN4 (Alomone) were used at dilutions of 1:200 and 1:400, respectively. Rabbit polyclonal antibodies against HCN3 (Alpha Diagnostic International, San Antonio, TX) were used at dilutions of 1:200 to 1:20. Contrary to the antibodies against HCN3 from Chemicon (Temecula, CA) and Alomone, this antibody does not show cross reactivity for hHCN1, hHCN2, or hHCN4 (Kouranova et al. 2008). Monoclonal antibodies against HCN2 and HCN3 from UC Davis/National Institute of Neurological Disorders and Stroke (NINDS)/National Institute of Mental Health (NIMH) NeuroMab Facility (Davis, CA) were used at 1:25. Mouse monoclonal and rabbit polyclonal antibodies against recombinant rat calretinin (Chemicon) were diluted at 1:1,000. Guinea-pig serum against VGLUT3 was kindly provided by Dr. Robert H. Edwards' laboratory (Department of Physiology, School of Medicine, University of California, San Francisco) and used at a 1:1,000 dilution. Alexa Fluor 488 F(ab′)2 fragment of goat anti-rabbit and Alexa Fluor 488 donkey anti-mouse IgG, Alexa Fluor 555 goat anti-mouse and Alexa Fluor 594 donkey anti-rabbit IgG, and Alexa Fluor 633 goat anti-guinea pig IgG (Molecular Probes, Eugene, OR) were used as secondary antibodies.

RESULTS

Voltage-gated ion channels in IHC afferent dendrites

To characterize voltage-gated conductances in IHC afferent dendrites, we used whole cell recordings from afferent dendrites directly where they contact the IHC. Recordings were performed in acutely excised apical cochlear turns from 7- to 14-day-old rats at room temperature. Resting membrane potentials of afferent dendrites were typically about −64 mV. We assume that the peripheral neurites of the recorded auditory nerve fibers were intact and connected with their spiral ganglion somata for the following reasons: First, in preparations where we separated the spiral ganglion from the cochlear coil, afferent dendrites were severely swollen and no recordings could be achieved. Second, in nine of nine experiments, where a fluorescent dye, Alexa Fluor 488 hydrazide salt (10 μM), had been included in the pipette solution and recordings had lasted >10 min, the unbranched afferent fibers could be traced back from the IHC toward the spiral ganglion for 200–500 μm (Supplemental Fig. S1)1 and, in two cases, the fluorescent marker had reached the spiral ganglion somata at a distance of 450–500 μm from the IHC.

Our study focused on the characterization of Ih. However, because voltage-gated conductances have not been described for IHC afferent dendrites, in the following paragraph we will briefly summarize the different conductances that were observed in response to voltage step protocols. From a holding potential of −84 mV, voltage steps between −104 and −4 mV were applied (Fig. 1A). Voltage-gated sodium currents were found in 114 of 155 afferent dendrite recordings. Sodium currents often “escaped” the voltage clamp (Fig. 1A, inset). This is not surprising because the recording site is at the very tip of the unmyelinated afferent fiber ending and the AP initiation site is most likely located further away along the myelinated peripheral process (Hossain et al. 2005; Lacas-Gervais et al. 2004; McLean et al. 2009). For this study sodium channels were not further investigated, but rather blocked with 1–2 μM TTX (Fig. 1B). Additionally, small calcium currents that could be blocked with 200 μM CdCl2 were detected in some recordings (data not shown). Because these small currents did not interfere with questions asked in this study, no effort was made to block them.

Fig. 1.

Fig. 1.

Current–voltage (I–V) relation recorded in an inner hair cell (IHC) afferent dendrite. Current responses to voltage steps in the absence (A) and the presence (B) of 1 μM tetrodotoxin (TTX). Voltage step protocol (inset in A): 200 ms voltage steps from −104 to −4 mV in 10 mV increments, from a holding potential of −84 mV. A: rapidly activating and inactivating sodium currents that sometimes escaped the voltage clamp (expanded trace shown in inset) were blocked by 1 μM TTX (B). C: current responses at 20 ms (solid circle) and 200 ms (open triangle) into the voltage steps after leak subtraction. A slowly activating inward current (i.e., a hyperpolarization-activated current [Ih]) was found at voltage steps to −94 mV or more negative potentials. Fast activating outward currents (within 5 ms) were activated at −64 mV and more positive voltages.

During depolarizing voltage steps, outward currents were observed in all 161 afferent dendrites recorded (Fig. 1). Outward currents reached their maximum within about 5 ms. The current–voltage (I–V) relations showed that outward currents activated at potentials as low as −64 mV (Fig. 1, B and C). We tested the effects of 4-AP (2–4 mM) and tetraethylammonium (TEA, 10–30 mM), drugs previously shown to inhibit the low-voltage activating potassium current (IKL) and the high-voltage activating potassium current (IKH), respectively, in spiral ganglion neurons (Szabo et al. 2002) and auditory brain stem neurons (Bal and Oertel 2001; Brew and Forsythe 1995; Cao et al. 2007; Manis and Marx 1991; Rathouz and Trussell 1998; Reyes et al. 1994; Rothman and Manis 2003a). 4-AP-sensitive and TEA-sensitive currents were observed in afferent dendrites and exhibited similar voltage dependent profiles to IKL and IKH, respectively. These conductances are still under investigation.

During hyperpolarizing voltage steps, a slowly developing inward current was found in 69 of 79 afferent recordings (Fig. 1, B and C). Voltage dependence and activation kinetics of this inward current were reminiscent of Ih recorded in dissociated spiral ganglion somata (Chen 1997; Mo and Davis 1997b).

Ih in afferent dendrites is mediated by HCN channels

To test whether Ih currents are mediated by HCN channels, we monitored Ih during the application of different blockers. In response to repeatedly applied negative voltage steps from −84 to −124 mV, an instantaneous inward current (Iinst) was followed by a slowly developing inward current (I0.5s) (Fig. 2C). Iinst is partially blocked by Ih blockers and consists of a mixture of Ih and other conductances (see Fig. 2, A–C) (Bal and Oertel 2000; Rodrigues and Oertel 2006). We therefore report only on the amplitude of the slowly developing component of Ih (I0.5sIinst). Ih currents showed some run-down during whole cell recordings and test protocols were designed accordingly. CsCl (2 mM) reversibly inhibited inward currents by 84 ± 11% (n = 8), whereas 2 mM BaCl2 caused a reversible inhibition by 14 ± 8% (n = 4) (Fig. 2, A, B, D, E, and F). This combination of a strong Cs+ block and a weak Ba2+ block has been described for HCN channels, whereas inward rectifier potassium channels typically show a substantial Ba2+ block (Kubo et al. 2005; Robinson and Siegelbaum 2003; van Welie et al. 2005). Additionally, 50 μM ZD7288, an antagonist of HCN channels (Shin et al. 2001), irreversibly inhibited Ih by 93 ± 13% (n = 4) (Fig. 2, C, E, and F).

Fig. 2.

Fig. 2.

Ih currents in afferent dendrites are mediated by hyperpolarization-activated cyclic nucleotide-gated cation (HCN) channels. A–C: in afferent dendrite recordings, hyperpolarizing voltage steps from −84 to −124 mV were applied every 10 s for 0.5 s (in 1 μM TTX). Current responses consisted of an instantaneous inward current (Iinst) (most obvious in C) and a slowly developing inward current (I0.5). A–C: representative traces before, during, and after application of 2 mM CsCl, 2 mM BaCl2, or 50 μM ZD7288 (4-ethylphenylamino-1,2-dimethyl-6-methylaminopyrimidinium chloride). Fast inward deflections during the recordings in A to B represent excitatory postsynaptic currents (EPSCs). D and E: diary plots of the Ih amplitude during application of drugs. Ih amplitude was measured as I0.5Iinst. The effects of CsCl and BaCl2 were mostly reversible; the effect of ZD7288 was irreversible. F: percentage reduction in Ih amplitude during CsCl, BaCl2, and ZD7288 application. The Ih amplitudes were determined from mean values of 5 consecutive traces in each condition. To compensate for rundown of Ih, current amplitudes in 2 mM CsCl or 2 mM BaCl2 were compared with mean value of respective control and recovery. The Ih amplitude in 50 μM ZD7288 was compared with control.

The reversal potential of Ih was estimated using the following protocol: different conditioning voltages were applied for 3 s (−124, −104, and −84 mV), followed by 10 ms voltage ramps (from −144 to −74 mV) (Fig. 3F). The initial current response (∼3 ms) during the voltage ramp was discarded due to contamination with uncompensated capacitance currents. The linear portions of the current responses (−120 to −74 mV) were extrapolated to the region where the responses intersect, corresponding to an estimated reversal potential of Ih at −45 ± 2 mV (n = 4, Fig. 3F). This reversal potential is consistent with a mixed cation channel permeable for sodium and potassium and is in the range of reversal potentials described for HCN channels (Bal and Oertel 2000; Banks et al. 1993; Cao et al. 2007; Chen 1997; Cuttle et al. 2001; Mo and Davis 1997a; Moosmang et al. 2001; Rodrigues and Oertel 2006; Santoro et al. 2000; van Welie et al. 2005). In summary, the pharmacological profile and reversal potential suggest that Ih in afferent dendrites is mediated by HCN channels.

Fig. 3.

Fig. 3.

Properties of Ih in afferent dendrites. A and B: Ih currents in response to voltage steps (every 10 s for 3 s, from a holding potential of −64 mV to voltages between −144 and −54 mV in 10 mV increments; see inset). External solution with: TTX (1–2 μM), 4-aminopyridine (4-AP, 2 mM), tetraethylammonium (TEA, 10–30 mM), and 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX, 10 μM). A: in control solution. B: with 200 μM cyclic adenosine monophosphate (cAMP) intracellularly and additionally 200 μM 8-bromoadenosine-3′,5′ cyclic monophosphate (8-Br-cAMP) extracellularly. C: I–V relations in control (n = 3, black) and with cAMP analogs (n = 4, red). D: voltage dependence of Ih measured from tail currents. Tail current amplitudes were normalized and fit with a Boltzmann equation. Vhalf and slope factors were −104 ± 3 mV, 11 ± 1 in control (n = 3, black), and −91 ± 2 mV, 11 ± 1 in cAMP analogs (n = 4, red). E: activation kinetics of Ih currents. Current responses to voltage steps from −134 to −104 mV were fit with 2 exponentials, providing 2 time constants (τfast, τslow). Both time constants were significantly faster for currents recorded with cAMP analogs (n = 4, red) compared with control (n = 3, black). F: reversal potential of Ih. Conditioning voltages were applied for 3 s (to −124, −104, or −84 mV), followed by 10 ms voltage ramps (from −144 to −74 mV) (top traces: voltage commands; bottom traces: current responses to the commanding voltage ramps). The reversal potential was −45.5 mV for this recording.

HCN channels in afferent dendrites are modulated by cAMP

The I–V relation of Ih was recorded while blocking voltage-gated sodium channels (1–2 μM TTX), potassium channels (2 mM 4-AP and 10–30 mM TEA), and AMPA receptors (10 μM CNQX) (Fig. 3, A–E). From a holding potential of −64 mV, voltage steps were applied for 3 s, from −144 to −54 mV in 10 mV increments, and were followed by a voltage step to −74 mV, during which tail currents were recorded (Fig. 3A). Ih was measured as I3sIinst. The I–V relation showed greater activation toward more negative voltages, with an Ih amplitude of 174 ± 48 pA at −144 mV (n = 3) (Fig. 3C, black trace). The voltage dependence of Ih was measured from tail currents and fitted by a Boltzmann equation (Fig. 3D, black traces). Half-maximum activation voltage (Vhalf) was −104 ± 3 mV and the slope factor was 11 ± 1 (n = 3). The activation curve shows that about 3% of Ih channels would be open at the resting membrane potential of the afferent dendrites (−65 mV). The activation range measured here corresponds to ranges measured under similar conditions for HCN channels in heterologous expression systems (Moosmang et al. 2001; Santoro et al. 2000) and Ih recorded in other neurons and spiral ganglion somata (Bal and Oertel 2000; Banks et al. 1993; Cao et al. 2007; Chen 1997; Cuttle et al. 2001; Mo and Davis 1997b; Rodrigues and Oertel 2006; van Welie et al. 2005).

Different HCN channel subunits activate on different timescales, with activation time constants ranging from milliseconds to seconds (Moosmang et al. 2001; Santoro et al. 2000). We measured the activation kinetics of Ih for current responses to voltage steps between −134 and −104 mV from a holding potential of −64 mV (Fig. 3E, black data points). The time course of activation was best fitted with two exponentials. At −134 mV, the time constants of the fast (τfast) and the slow (τslow) component were 66 ± 13 ms (n = 3) and 929 ± 40 ms, respectively [Afast/(Afast + Aslow) = 0.83 ± 0.09]. At more positive potentials, activation slowed down for both components.

It is well known that intracellular cyclic nucleotides can modulate HCN channel activity (Moosmang et al. 2001; Santoro et al. 2000; Wainger et al. 2001). Typically, with increased cAMP levels, the activation curve is shifted to more positive values and the activation kinetics speeds up. We therefore measured the I–V relation in the presence of cAMP analogs (Fig. 3B). cAMP (200 μM) was added to the pipette solution and, additionally, 200 μM of the membrane permeable 8-Br-cAMP (8-bromoadenosine-3′,5′ cyclic monophosphate) was added to the external solution. In the presence of cAMP analogs, the current amplitude of Ih increased significantly compared with control (from 174 ± 48 pA in control to 332 ± 91 pA in cAMP; at −144 mV, n = 4, P < 0.05 for all voltages tested) (Fig. 3, B and C, red traces). The activation curve shifted to more positive values, by about 12 mV (Fig. 3D, red traces), with Vhalf at −91 ± 2 mV (n = 4, P < 0.05) and no significant change in the slope factor (11 ± 1, n = 4, P = 0.611). Under these conditions, at −65 mV, about 9% of Ih would be active and have an estimated conductance of 1.4 nS (332 pA × 0.085/20 mV; reversal potential −45 mV). In the presence of cAMP analogs, activation time constants were significantly faster with a τfast of 23 ± 6 ms and a τslow of 175 ± 76 ms at a holding potential of −134 mV [n = 4, P < 0.05 for all voltages tested: Afast/(Afast + Aslow) = 0.85 ± 0.05] (Fig. 3E, red data points).

HCN channel subunit expression in IHC afferent dendrites

To investigate the expression pattern of HCN channel subunits in IHC afferent dendrites, we performed immunolabeling experiments with antibodies raised against the four known HCN channel subunits HCN1–HCN4 (Moosmang et al. 2001; Robinson and Siegelbaum 2003). We examined HCN distribution in the apical part of the cochlea in prehearing animals, at P9–P10, to match our recordings from afferent dendrites, and at P20–P21, to analyze HCN subunit expression in hearing animals. No immunolabeling of HCN3 above background was found and therefore the result could not be interpreted. Figure 4 shows overviews of apical cochlear turns at P9–P10 labeled for HCN1 or HCN4. At this age, we did not find HCN2 labeling. For better orientation, IHCs were labeled for the vesicular glutamate transporter VGLUT3 (Seal et al. 2008) (Fig. 4B). HCN1 and HCN4 specific labeling in the organ of Corti was found in the area directly below the IHCs, suggestive of a labeling of the unmyelinated peripheral processes of the afferent neurons. Additionally, HCN1 and HCN4 immunoreactivity was found in most of the spiral ganglion somata (Fig. 4 and Supplemental Fig. S2). The labeling was concentrated at the plasma membrane of the somata and in their processes within the ganglion close to the somata (Supplemental Fig. S2). At P20–P21, similar to the labeling pattern at P9–P10, labeling for HCN1, HCN4, and additionally for HCN2 was found directly below the IHCs (Fig. 5) and in the spiral ganglion somata (Supplemental Fig. S2).

Fig. 4.

Fig. 4.

HCN subunit expression pattern in the rat cochlea at P9. Three-dimensional reconstruction of confocal images from cochlear whole-mount preparations, apical turns. A: HCN1 labeling (green). B: HCN4 labeling (green). Vesicular glutamate transporter VGLUT3 (red) was used as a marker for IHCs. HCN1 and HCN4 labeling was found in the inner spiral plexus (isp) under the row of IHCs (ihc, arrowhead) as well as in the somata of spiral ganglion neurons (sgn). Scale bars: 50 μm.

Fig. 5.

Fig. 5.

HCN subunits are localized in afferent dendrites. A: 3-dimensional reconstruction of calretinin and HCN1-labeled whole-mount rat organ of Corti preparation, apical turn at P9. HCN1 labeling (green) is concentrated in the basolateral region of the IHCs. Calretinin (red) labels IHCs (ihc) and afferent dendrites. B–D: close-up view showing single confocal laser-scanning micrographs. As seen in the merged view (D), HCN1 (B) and calretinin (C) immunolabeling overlap in some afferent dendrites (arrowheads). E–M: single confocal laser-scanning micrographs of whole-mount organs of Corti preparations, apical turns at P21. Preparations were colabeled for HCN1, HCN2, or HCN4 (green) and calretinin (red). Arrowheads indicate examples of double-labeled afferent dendrites. Note examples of ringlike HCN labeling surrounding calretinin labeling (open arrowheads). Some fibers were labeled for HCN but not for calretinin (asterisks). Scale bars: 5 μm.

IHCs are surrounded by different cell types, including afferent fibers, efferent fibers, and supporting cells. To confirm that the afferent dendrites in the IHC area express the different HCN subunits, as indicated by our electrophysiological recordings, we performed double immunolabeling experiments for calretinin and HCN channel subunits. Calretinin is a calcium-binding protein involved in calcium buffering and transport. In the rat cochlea, calretinin has been detected both in the cytoplasm of IHCs and in most spiral ganglion neurons including their afferent dendrites (Dechesne et al. 1991, 1993). High resolution confocal imaging confirmed the coexpression of the HCN subunits and calretinin in the same afferent dendrites at P9–P10 and P20–P21 (Fig. 5). When assessing the labeling close to the IHCs at P9–P10, some afferent dendrites could be identified by their strong labeling with calretinin and were also found positive for HCN1 (Fig. 5, A–D, arrowheads) or HCN4 (data not shown). At P20–P21, HCN immunolabeling was more sharply defined, most likely reflecting a higher concentration or more intense clustering of the HCN channel subunits in afferent dendrites in the more mature organ of Corti. Most afferent dendrites positive for HCN1, HCN2, and HCN4 were also positive for calretinin labeling, indicating that HCN1, HCN2, and HCN4 are indeed localized in the afferent dendrites (Fig. 5, E–M, arrowheads). In some dendrites, HCN labeling appeared to surround the calretinin labeling in a ringlike fashion (open arrowheads), consistent with a clustering of the HCN channels in the plasma membrane. Some HCN positive dendrites were not calretinin positive (Fig. 5, asterisks). Similarly, not all HCN positive spiral ganglion somata were positive for calretinin labeling (data not shown). We conclude that the HCN subunits 1, 2, and 4 are expressed in afferent dendrites at the IHC afferent synapse.

EPSP waveforms in IHC afferent dendrites

In the CNS, dendritic Ih has been shown to shape EPSPs and to modulate excitability in neurons (Magee 2000). To investigate whether Ih affects the EPSP waveform in IHC afferent dendrites, we first characterized the EPSP waveform at rest. To isolate synaptic activity and block the generation of APs, experiments were performed in 1 μM TTX. Resting membrane potentials of afferent dendrites were −64 ± 8 mV (n = 34) when recorded in our standard extracellular solution containing 5.8 mM K+ and 1.3 mM Ca2+. Synaptic events occurred at a rate of 0.7 ± 0.9/s (n = 47; 13,575 synaptic events analyzed).

As shown before for EPSCs (Glowatzki and Fuchs 2002), some EPSPs appeared “multiphasic,” composed of multiple overlapping events. Other EPSPs appeared “monophasic,” presenting a monoexponential decay (Fig. 6F). To investigate rise and decay times, only monophasic events were analyzed. In three recordings, for direct comparison, both EPSCs and EPSPs were characterized in the same recording (Fig. 6, A–K). EPSCs were recorded at a holding potential of −94 mV and EPSPs at the resting membrane potential of the afferent dendrite in 5.8 mM K+. EPSPs were about three to four times slower than EPSCs, with a 10–90% rise time of 1.14 ± 0.17 ms compared with 0.36 ± 0.09 ms and a time constant of decay of 5.13 ± 1.69 ms compared with 1.16 ± 0.13 ms at room temperature (n = 3, 381 EPSPs; n = 8, 1,807 EPSCs analyzed) (Fig. 6, C, D, G, and H). For EPSPs, we found a wide range of amplitudes in every recording, varying from 1 to 35 mV; the shape of the amplitude distributions also varied widely. The amplitude distributions of two fibers were highly skewed, with median EPSP amplitudes of 2.3 and 2.4 mV, and most of the EPSPs were <10 mV (Fig. 6, I and J). For a third fiber, EPSP amplitudes spread more evenly within a range between 1 and 35 mV, with a median EPSP amplitude of 13.8 mV (Fig. 6K). However, for both EPSCs and EPSPs, rise and decay times did not change much over this wide range of amplitudes (Fig. 6, C, D, G, and H).

Fig. 6.

Fig. 6.

EPSCs and excitatory postsynaptic potentials (EPSPs) recorded at the IHC afferent synapse. A–K: whole cell recording from an afferent dendrite in the presence of 1 μM TTX showing EPSCs (A, B) (holding potential −94 mV) and EPSPs (E, F). B and F: overlaid representative traces of monophasic EPSCs and EPSPs on an expanded timescale. C, D, G, and H: 10–90% rise time (rise) or decay time constants (τdecay) plotted against the EPSC or EPSP amplitude. Rise and τdecay for EPSCs were 0.33 ± 0.14 and 1.24 ± 0.20 ms (324 EPSCs analyzed). Rise and τdecay for EPSPs were 0.96 ± 0.12 and 3.81 ± 0.36 ms (241 EPSPs analyzed). EPSP waveforms remained relatively invariable over the wide range of EPSP amplitudes. I–K: EPSP amplitude distributions (bin size 1 mV) from 3 afferent dendrite recordings. Median EPSP amplitudes were 2.3, 2.4, and 13.8 mV and resting membrane potentials were −75, −56, and −68 mV, respectively. The number of events analyzed is indicated in each panel. L–N: whole cell current-clamp recording in the absence of TTX. A mixture of EPSPs and spikes was observed. M: overlaid representative traces of spikes on an expanded timescale. Spike threshold (arrow) was −47 mV for this recording. N: amplitude distribution (bin size 1 mV). A wide gap in amplitude histogram distinguishes spikes from EPSPs.

In auditory nerve fibers, the spike initiation zone is believed to be located close to the IHC afferent synapse, on the peripheral process (Hossain et al. 2005; Robertson 1976). A heminode with a high expression level of sodium channels, is <50 μm away from the afferent synapse (Hossain et al. 2005; McLean et al. 2009). The expression of sodium channels was also reported on the unmyelinated ending of the afferent fiber, peripheral to the heminode (Hossain et al. 2005). Indeed, when recording in the absence of TTX, at a resting membrane potential of −67 ± 5 mV (n = 7), 18 ± 22% of the events were spikes rather than EPSPs (range 0.5–50.5%, n = 4 afferent dendrites with >100 analyzable events each; total number of events analyzed: 829) (Fig. 6, L–N). Spikes were discriminated from EPSPs by their large and uniform amplitudes (39 ± 7 mV, n = 7, 171 spikes) that appeared as a separate group of events in the amplitude histograms (Fig. 6N). Additionally, about half of the spikes exhibited sudden slope changes during their rise (Fig. 6M, arrow), suggestive of a spike threshold at −50 ± 4 mV (n = 7, 97 spikes). The 10–90% rise time of the spikes (including the rise of the initiating EPSPs) was 2.29 ± 0.75 ms and the spike half-width was 2.83 ± 0.86 ms (n = 7, 171 spikes).

HCN channel activity shortens the EPSP waveform in IHC afferent dendrites

Next, we tested whether Ih is involved in shaping the EPSP waveform. To block Ih, we first applied CsCl, to allow for recovery of a possible effect. Without any cAMP analog added, 2 mM CsCl significantly increased τdecay of the EPSP by 27 ± 17% (from 6.37 ± 0.64 to 8.04 ± 1.02 ms and, after washout, to 6.98 ± 0.53 ms; n = 4 fibers, 617 EPSPs analyzed, P < 0.05). Neither mean EPSP amplitude (control vs. CsCl: 13 ± 6 vs. 14 ± 8 mV, P = 0.839) nor 10–90% rise time (1.31 ± 0.20 vs. 1.48 ± 0.20 ms, P = 0.146) changed significantly. In the presence of cAMP analogs (200 μM cAMP intracellularly and, additionally, 200 μM 8-Br-cAMP extracellularly), application of CsCl increased τdecay by 37 ± 31% (from 5.25 ± 1.37 to 7.15 ± 1.29 ms, n = 7, 878 EPSPs analyzed, P < 0.05) (Fig. 7, A–C). Again, EPSP mean amplitude (control vs. CsCl: 10 ± 4 vs. 11 ± 6 mV, P = 0.337) and 10–90% rise time (1.23 ± 0.20 vs. 1.31 ± 0.30 ms, P = 0.376) did not significantly change. Application of CsCl hyperpolarized the resting membrane potential of the afferent dendrite by about 1 mV (from −64 ± 3 to −65 ± 3 mV, n = 11, P = 0.002).

Fig. 7.

Fig. 7.

Ih shortens EPSPs in afferent dendrites. A–H: whole cell current-clamp recording from afferent dendrites. Recordings were done in the presence of cAMP analogs (200 μM 8-Br-cAMP extracellularly and additionally 200 μM cAMP intracellularly). A–C: EPSP waveform before and while blocking Ih with 2 mM CsCl. A: average EPSP waveform before (black) and during application of 2 mM CsCl (red). B: EPSP decay time constants (τdecay) plotted against EPSP amplitudes before and while blocking Ih with 2 mM CsCl; control: τdecay = 4.97 ms (black, 34 EPSPs); in 2 mM CsCl: τdecay = 7.18 ms (red, 41 EPSPs). C: summarized results from 7 recordings. D–F: EPSP waveform before and during application of 50 μM ZD7288. D: average EPSP waveform before (black) and during application of 50 μM ZD7288 (magenta). E: EPSP decay time constants (τdecay) plotted against EPSP amplitudes; control: τdecay = 5.38 ms (black, 22 EPSPs), in 50 μM ZD7288: τdecay = 8.30 ms (magenta, 16 EPSPs). F: summarized results from 6 recordings.

Immature IHCs express inward rectifier potassium channels that can be inhibited by extracellular Cs+ (Marcotti et al. 2003). Indeed, the frequency of synaptic events was higher in CsCl (control vs. CsCl: 1.01 ± 1.66 vs. 2.23 ± 2.44 events/s, n = 16, 4,357 EPSP or EPSCs analyzed, P = 0.025). To exclude the possibility that CsCl by some presynaptic effect changes the EPSC waveform and therefore affects the EPSP waveform, we tested the effect of CsCl on EPSCs and found no significant difference compared with control (control vs. CsCl: EPSC mean amplitude: 159 ± 91 vs. 168 ± 99 pA, P = 0.838; 10–90% rise time: 0.38 ± 0.14 vs. 0.38 ± 0.14 ms, P = 0.796; τdecay: 1.24 ± 0.40 vs. 1.43 ± 0.57 ms, P = 0.383; n = 3, 626 EPSCs analyzed). These data suggest that changes in the EPSP waveform during CsCl application are due to its effect on the afferent dendrite.

Application of the irreversible HCN channel blocker ZD7288 exhibited greater effects than those of CsCl on the EPSP waveform, without changing the EPSP frequency (control vs. ZD7288: 0.73 ± 0.76 vs. 0.70 ± 0.71 events/s, P = 0.582) (Fig. 7, D–F). During application of 50 μM ZD7288 (in the presence of 200 μM cAMP intracellularly and, additionally, 200 μM 8-Br-cAMP extracellularly), τdecay increased by 47 ± 24% (from 4.22 ± 0.98 to 6.19 ± 1.67 ms, n = 6, 593 EPSPs analyzed, P = 0.002) and the resting membrane potential was hyperpolarized by about 4 mV (from −63 ± 6 to −67 ± 7 mV, P < 0.001, n = 11). EPSP amplitude (8 ± 3 vs. 9 ± 4 mV, P = 0.378) and 10–90% rise time (1.40 ± 0.32 vs. 1.49 ± 0.31 ms, P = 0.082) did not significantly change. The Ih-induced hyperpolarization of the membrane potential could change the activity of additional ion channels that also might affect the EPSP waveform as it has been shown in a computational model of cochlear nucleus neurons (Rothman and Manis 2003b) or in recordings from dendrites of both hippocampal neurons (George et al. 2009) and frontal cortex pyramidal neurons (Day et al. 2005). Therefore to exclude possible effects on the EPSP waveform by hyperpolarization, membrane potentials were manually reset during ZD7288 application to their control values by constant current injection in three recordings. A significant increase of 30 ± 11% in τdecay compared with control was still observed (Fig. 7F, τdecay in control: 4.22 ± 0.98 ms; τdecay in ZD7288: 5.25 ± 1.51 ms, P = 0.03). This result indicates that Ih shapes the EPSP waveform directly by contributing to the membrane resting conductance. The application not only of CsCl but also of ZD7288 increased τdecay over the entire range of EPSP amplitudes (Fig. 7, B and E). We conclude that the activity of Ih in afferent dendrites shortens the EPSP waveform over the whole range of EPSP amplitudes.

Application of ZD7288 also caused minor changes in the spike waveform. The spike amplitude increased by 9 ± 4% (from 32 ± 5 to 35 ± 4 mV, P = 0.026; n = 3, 155 spikes analyzed) and spike half-width increased by 14 ± 8% (from 3.05 ± 0.53 to 3.45 ± 0.39 ms, P = 0.046). The 10–90% rise time did not change significantly (control vs. ZD7288: 1.24 ± 0.20 vs. 1.29 ± 0.19 ms, P = 0.757).

DISCUSSION

Properties of Ih in IHC afferent dendrites

Dendritic Ih has been shown to play an important role in setting firing rates in nerve fibers (Magee and Johnston 2005). Here we have used whole cell recordings to provide an initial characterization of Ih present in postnatal IHC afferent dendrites. The pharmacological profile of Ih, its reversal potential, its sensitivity to cAMP, and the expression of the HCN subunits 1, 2, and 4 in afferent dendrites indicate that Ih is mediated by HCN channels. In immunolabeling experiments we found HCN1 and HCN4 in postnatal afferent fibers and HCN1, HCN2, and HCN4 after hearing onset. No immunolabeling of HCN3 above background was found in either the spiral ganglion or the afferent dendrites at both ages and thus we cannot make a statement about the expression of HCN3. In adult guinea pig spiral ganglion somata, the expression of all four HCN subunits has been reported (Bakondi et al. 2009). The different result regarding HCN3 expression might be due to a difference in the species or to the different antibodies used. No specific HCN1 or HCN2 labeling has been found in mouse cochlear hair cells and mice lacking HCN1, HCN2, or both exhibited normal transduction currents (Horwitz et al. 2010).

We found each HCN subunit in a high percentage of afferent dendrites. It is therefore highly likely that heteromeric channels are formed. However, it is difficult to unequivocally deduce the composition of HCN channels in afferent dendrites from the properties of the recorded Ih. When measured in the same experimental conditions, homomeric HCN1 channels activate at more positive voltages and show faster kinetics than homomeric HCN2 and HCN4 channels. On the other hand, HCN1 shows the least sensitivity to cyclic nucleotides, followed by HCN2 to HCN4 (Moosmang et al. 2001; Santoro et al. 2000). Heteromeric HCN channels have been shown to exhibit intermediate properties compared with those of homomeric channels (Chen et al. 2001; Ulens and Tytgat 2001). Additionally, HCN channels with the same subunit composition exhibit variable characteristics when measured in different expression systems or tissues (Moosmang et al. 2001; Santoro et al. 2000; Wainger et al. 2001). Knock-out animals for HCN1, HCN2, and HCN4 do exist, although reports on their behavior do not mention any test of auditory function (Harzheim et al. 2008; Herrmann et al. 2007; Ludwig et al. 2003; Nolan et al. 2003, 2004; Stieber et al. 2003).

A range of activation voltages and cyclic nucleotide sensitivities has been described for Ih in studies on auditory neurons. For example, in bushy cells, a Vhalf of −94 to −84 mV and positive shift of Vhalf with cAMP were found (Cuttle et al. 2001; Leao K et al. 2006), whereas in octopus cells, a Vhalf of −65 mV and no shift with cAMP were reported (Bal and Oertel 2000). The properties we found for Ih in afferent dendrites are comparable to those found for Ih in spiral ganglion somata. In afferent dendrites, Vhalf was −104 mV, similar to Vhalf reported for spiral ganglion somata (−101 mV; Chen 1997; −78 to −122 mV; Mo and Davis 1997b). Both in spiral ganglion somata and afferent dendrites, Vhalf shifted positively in the presence of cAMP analogs.

In control conditions, we found 3% and, in the presence of cAMP analogs, 9% of Ih open at rest. The experimental conditions might underestimate the size of Ih in vivo; we recorded at room temperature and the amplitude of Ih has been found to be temperature sensitive in auditory neurons (Cao and Oertel 2005; Cao et al. 2007; Rodrigues and Oertel 2006). We recorded in immature cochleae, however, our immunolabeling experiments suggest a maturation of HCN channel subunit expression in hearing animals. For example, HCN2 labeling of afferent dendrites was found in 3-wk-old cochleae but not before hearing onset. HCN2 has a higher sensitivity to cAMP compared with that of HCN1 and thus properties of Ih may change with maturation. In our whole cell recordings, second messengers may be subject to wash-out and therefore the Vhalf measured may be more negative compared with that under in vivo conditions. Additionally, under in vivo conditions, input from lateral efferent fibers may up-regulate intradendritic cAMP levels (see last paragraph) and in vitro transmitter release from these fibers may be absent or abnormal.

Role of Ih in auditory neurons

Multiple studies have shown that auditory neurons involved in processing timing information express Ih. Examples include bushy cells (Cao et al. 2007; Leao K et al. 2006; Leao R et al. 2005; Rothman and Manis 2003a), octopus cells (Bal and Oertel 2000; Cao and Oertel 2005; Golding et al. 1995, 1999), and stellate cells (Rodrigues and Oertel 2006; Rothman and Manis 2003b) of the cochlear nucleus, medial superior olive neurons (Scott et al. 2005), nucleus laminaris neurons (Kuba et al. 2005; Yamada et al. 2005), lateral superior olive neurons (Barnes-Davies et al. 2004; Leao K et al. 2006), medial nucleus of the trapezoid body (MNTB) neurons (Banks et al. 1993; Cuttle et al. 2001; Leao K et al. 2006), the calyx of Held (Cuttle et al. 2001), and inferior colliculus neurons (Koch and Grothe 2003). Additionally, Ih was found in the somata of spiral ganglion neurons (Chen 1997; Mo and Davis 1997b). Ih can be partially active at rest, contributing to a lower membrane input resistance and thus shortening EPSPs. Shorter EPSPs reduce the time window of temporal summation and therefore improve coincidence detection (Golding et al. 1995; Koch and Grothe 2003; Yamada et al. 2005). Similarly, Ih expressed in dendrites of nonauditory CNS neurons shortened EPSPs and thereby reduced summation of synaptic activity (Day et al. 2005; Magee 2000; Magee and Johnston 2005). The nonuniform distribution of HCN channels along dendrites compartmentalized distal dendrites from the somata (Berger et al. 2003), limiting action potential (AP) back-propagation, and thereby decreasing hyperexcitability (Tsay et al. 2007; Ying et al. 2007). In motion-sensitive neurons in the superior colliculus (Endo et al. 2008), dendritic Ih was shown to keep AP timing at a minimum jitter and with short latencies. It is not surprising that similar mechanisms might be used at the IHC afferent synapse, as discussed in the following text.

Properties of EPSPs in afferent fibers and effects of Ih on EPSPs

Here we describe EPSP waveforms at the IHC afferent synapse. EPSPs have 10–90% rise times of about 1 ms and time constants of decay of about 5 ms. In comparison, the duration of EPSPs recorded in vivo in guinea pig (Palmer and Russell 1986; Siegel 1992; Siegel and Dallos 1986) was about 1–2 ms. The slower kinetics of EPSPs reported here could be due to the immature age of the recorded dendrites. AMPA mediated EPSCs at IHC afferent synapses speed up by a factor of 2, in a comparison of the immature age used in this study with 3-wk-old animals (Grant et al. 2010). Also, our recordings were performed at room temperature and AMPA receptor kinetics are temperature sensitive, with a Q10 of >2 (Postlethwaite et al. 2007). The amplitudes and gating kinetics of voltage-gated ion channels, such as Ih and IKL that are known to shorten EPSPs and APs in auditory neurons, are generally larger and faster in posthearing animals (Scott et al. 2005) and at body temperature (Cao and Oertel 2005; Cao et al. 2007; Rodrigues and Oertel 2006).

As shown for EPSCs (Glowatzki and Fuchs 2002; Grant et al. 2010), EPSP amplitudes covered an impressively wide range, between 1 and 35 mV. The median EPSP amplitudes varied widely between fibers, between 2 and 14 mV. Spikes recorded in afferent dendrites had a threshold at about −50 mV and roughly 20% of EPSPs activated spikes. Therefore for small EPSPs, postsynaptic summation of EPSPs might be necessary to reach the threshold for AP generation. In simultaneous recordings from hair cells and afferent dendrites, summation of EPSCs was observed during the early response to step depolarizations of the IHC membrane potential (Goutman and Glowatzki 2007). Our results showed that EPSPs slowed down by 47% when Ih was blocked by ZD7288. These data suggest that depending on the level of Ih active, the time window of EPSP summation may vary, allowing for regulation of firing rate in auditory nerve fibers.

Block of Ih with ZD7288 also induced a hyperpolarization of the membrane potential by about 4 mV. When the membrane potential was reset to control values during block of Ih, EPSPs still slowed down by roughly 30%. This result indicates that Ih shapes the EPSP waveform directly by contributing to the resting membrane conductance. Ih may also act indirectly on the EPSP waveform, by depolarizing the membrane potential and thereby activating other ion channels such as IKL that might additionally change the resting membrane conductance (Rothman and Manis 2003b). However, our data set here does not prove or reject this scenario because the changes in EPSP waveform during block of Ih with and without hyperpolarization were not significantly different.

During block of Ih, and with a change in the resting membrane conductance, we expected to see not only slowing of the EPSP waveform but also an increase in the EPSP amplitude (Magee 1998). The EPSP amplitude increased slightly, although not significantly. This is most likely due to the wide range of EPSP amplitudes in individual recordings. We suspect that if a larger data set was available, the difference in amplitude might have reached statistical significance. The fact that there was a small but significant increase in the spike amplitude (representing a less variable waveform compared with the EPSPs) further supports this idea.

HCN channels in afferent dendrites as possible targets for lateral efferent transmission

IHC afferent dendrites receive efferent innervation from the lateral superior olivary complex. Multiple transmitters such as acetylcholine, dopamine, γ-aminobutyric acid, and opiate peptides have been found in lateral efferent terminals (Eybalin 1993). Lesion of lateral efferent nerves disrupted temporal coding and increased susceptibility to acute acoustic trauma (Darrow et al. 2006, 2007). Intracochlear perfusion of putative lateral efferent neurotransmitters like dopamine affected firing rates in the auditory nerve (Ruel et al. 2001). However, the cellular mechanisms underlying these modulatory effects are unclear. One possibility is that lateral efferent transmitters modulate dendritic ion channels such as Ih, for example via second messenger cascades, thereby subsequently affecting auditory nerve firing. Indeed, such modulation of Ih by a neurotransmitter has been demonstrated in auditory neurons. In rat MNTB neurons, noradrenaline and cAMP analogs increased the amplitude and shifted the activation curve of Ih (Banks et al. 1993). In chick nucleus laminaris, noradrenaline enhanced temporal precision of EPSPs by regulating Ih activity in the postsynapse (Yamada et al. 2005). In addition to cyclic nucleotides, phosphorylation by protein kinases, pH, and lipid second messengers have been shown to modulate HCN channel activity (Fogle et al. 2007; Pian et al. 2007; Robinson and Siegelbaum 2003; Zolles et al. 2006). Our finding that a cAMP-sensitive Ih in afferent dendrites modulates the EPSP waveform supports the idea that modulation of temporal coding could occur directly at the first synapse of the auditory pathway, possibly via lateral efferent inputs.

DISCLOSURES

No conflicts of interest are declared by the authors.

Supplementary Material

[Supplemental Figures]
00506.2009_index.html (822B, html)

ACKNOWLEDGMENTS

We thank R. H. Edwards' laboratory for kindly providing the antibodies against VGLUT3, J. Gibas for excellent technical assistance with confocal microscopy, and L. Grant for comments on the manuscript.

Monoclonal antibodies against HCN2 and HCN3 were obtained from UC Davis/National Institute of Neurological Disorders and Stroke (NINDS)/National Institute of Mental Health (NIMH) NeuroMab Facility, supported by NINDS Grant U24-NS-050606 and maintained by the Department of Pharmacology, School of Medicine, University of California, Davis. This work was supported by National Institute on Deafness and Other Communication Disorders (NIDCD) Grants DC-006476 to E. Glowatzki and DC-008860 to D. Bergles, a research grant from the Deafness Research Foundation to E. Yi, a European Molecular Biology Organization Fellowship ALTF 952-2006 to I. Roux, National Institute of Diabetes and Digestive and Kidney Diseases Grant R24-DK-064388 to the Ross Confocal Facility, and NIDCD Grant P30 DC-005211 to the Center for Hearing and Balance Histology Core.

Footnotes

1

The online version of this article contains supplemental data.

REFERENCES

  1. Bakondi G, Por A, Kovacs I, Szucs G, Rusznak Z. Hyperpolarization-activated, cyclic nucleotide-gated, cation non-selective channel subunit expression pattern of guinea-pig spiral ganglion cells. Neuroscience 158: 1469–1477, 2009. [DOI] [PubMed] [Google Scholar]
  2. Bal R, Oertel D. Hyperpolarization-activated, mixed-cation current (Ih) in octopus cells of the mammalian cochlear nucleus. J Neurophysiol 84: 806–817, 2000. [DOI] [PubMed] [Google Scholar]
  3. Bal R, Oertel D. Potassium currents in octopus cells of the mammalian cochlear nucleus. J Neurophysiol 86: 2299–2311, 2001. [DOI] [PubMed] [Google Scholar]
  4. Banks MI, Pearce RA, Smith PH. Hyperpolarization-activated cation current (Ih) in neurons of the medial nucleus of the trapezoid body: voltage-clamp analysis and enhancement by norepinephrine and cAMP suggest a modulatory mechanism in the auditory brain stem. J Neurophysiol 70: 1420–1432, 1993. [DOI] [PubMed] [Google Scholar]
  5. Banks MI, Smith PH. Intracellular recordings from neurobiotin-labeled cells in brain slices of the rat medial nucleus of the trapezoid body. J Neurosci 12: 2819–2837, 1992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Barnes-Davies M, Barker MC, Osmani F, Forsythe ID. Kv1 currents mediate a gradient of principal neuron excitability across the tonotopic axis in the rat lateral superior olive. Eur J Neurosci 19: 325–333, 2004. [DOI] [PubMed] [Google Scholar]
  7. Berger T, Senn W, Luscher HR. Hyperpolarization-activated current Ih disconnects somatic and dendritic spike initiation zones in layer V pyramidal neurons. J Neurophysiol 90: 2428–2437, 2003. [DOI] [PubMed] [Google Scholar]
  8. Brew HM, Forsythe ID. Two voltage-dependent K+ conductances with complementary functions in postsynaptic integration at a central auditory synapse. J Neurosci 15: 8011–8022, 1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  9. Cao XJ, Oertel D. Temperature affects voltage-sensitive conductances differentially in octopus cells of the mammalian cochlear nucleus. J Neurophysiol 94: 821–832, 2005. [DOI] [PubMed] [Google Scholar]
  10. Cao XJ, Shatadal S, Oertel D. Voltage-sensitive conductances of bushy cells of the mammalian ventral cochlear nucleus. J Neurophysiol 97: 3961–3975, 2007. [DOI] [PubMed] [Google Scholar]
  11. Chen C. Hyperpolarization-activated current (Ih) in primary auditory neurons. Hear Res 110: 179–190, 1997. [DOI] [PubMed] [Google Scholar]
  12. Chen S, Wang J, Siegelbaum SA. Properties of hyperpolarization-activated pacemaker current defined by coassembly of HCN1 and HCN2 subunits and basal modulation by cyclic nucleotide. J Gen Physiol 117: 491–504, 2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Cuttle MF, Rusznak Z, Wong AY, Owens S, Forsythe ID. Modulation of a presynaptic hyperpolarization-activated cationic current (Ih) at an excitatory synaptic terminal in the rat auditory brainstem. J Physiol 534: 733–744, 2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  14. Darrow KN, Maison SF, Liberman MC. Cochlear efferent feedback balances interaural sensitivity. Nat Neurosci 9: 1474–1476, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Darrow KN, Maison SF, Liberman MC. Selective removal of lateral olivocochlear efferents increases vulnerability to acute acoustic injury. J Neurophysiol 97: 1775–1785, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Day M, Carr DB, Ulrich S, Ilijic E, Tkatch T, Surmeier DJ. Dendritic excitability of mouse frontal cortex pyramidal neurons is shaped by the interaction among HCN, Kir2, and Kleak channels. J Neurosci 25: 8776–8787, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Dechesne CJ, Winsky L, Kim HN, Goping G, Vu TD, Wenthold RJ, Jacobowitz DM. Identification and ultrastructural localization of a calretinin-like calcium-binding protein (protein 10) in the guinea pig and rat inner ear. Brain Res 560: 139–148, 1991. [DOI] [PubMed] [Google Scholar]
  18. Dechesne CJ, Winsky L, Moniot B, Raymond J. Localization of calretinin mRNA in rat and guinea pig inner ear by in situ hybridization using radioactive and non-radioactive probes. Hear Res 69: 91–97, 1993. [DOI] [PubMed] [Google Scholar]
  19. Endo T, Tarusawa E, Notomi T, Kaneda K, Hirabayashi M, Shigemoto R, Isa T. Dendritic Ih ensures high-fidelity dendritic spike responses of motion-sensitive neurons in rat superior colliculus. J Neurophysiol 99: 2066–2076, 2008. [DOI] [PubMed] [Google Scholar]
  20. Fogle KJ, Lyashchenko AK, Turbendian HK, Tibbs GR. HCN pacemaker channel activation is controlled by acidic lipids downstream of diacylglycerol kinase and phospholipase A2. J Neurosci 27: 2802–2814, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Fuchs PA. Time and intensity coding at the hair cell's ribbon synapse. J Physiol 566: 7–12, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. Funabiki K, Koyano K, Ohmori H. The role of GABAergic inputs for coincidence detection in the neurones of nucleus laminaris of the chick. J Physiol 508: 851–869, 1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Gardner SM, Trussell LO, Oertel D. Time course and permeation of synaptic AMPA receptors in cochlear nuclear neurons correlate with input. J Neurosci 19: 8721–8729, 1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. Gardner SM, Trussell LO, Oertel D. Correlation of AMPA receptor subunit composition with synaptic input in the mammalian cochlear nuclei. J Neurosci 21: 7428–7437, 2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. George MS, Abbott LF, Siegelbaum SA. HCN hyperpolarization-activated cation channels inhibit EPSPs by interactions with M-type K(+) channels. Nat Neurosci 12: 577–584, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  26. Glowatzki E, Fuchs PA. Transmitter release at the hair cell ribbon synapse. Nat Neurosci 5: 147–154, 2002. [DOI] [PubMed] [Google Scholar]
  27. Glowatzki E, Grant L, Fuchs P. Hair cell afferent synapses. Curr Opin Neurobiol 18: 389–395, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Golding NL, Ferragamo MJ, Oertel D. Role of intrinsic conductances underlying responses to transients in octopus cells of the cochlear nucleus. J Neurosci 19: 2897–2905, 1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Golding NL, Robertson D, Oertel D. Recordings from slices indicate that octopus cells of the cochlear nucleus detect coincident firing of auditory nerve fibers with temporal precision. J Neurosci 15: 3138–3153, 1995. [DOI] [PMC free article] [PubMed] [Google Scholar]
  30. Goutman JD, Glowatzki E. Time course and calcium dependence of transmitter release at a single ribbon synapse. Proc Natl Acad Sci USA 104: 16341–16346, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  31. Grant L, Yi E, Glowatzki E. Two modes of release shape the postsynaptic response at the inner hair cell ribbon synapse. J Neurosci 30: 4210–4220, 2010. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Harzheim D, Pfeiffer KH, Fabritz L, Kremmer E, Buch T, Waisman A, Kirchhof P, Kaupp UB, Seifert R. Cardiac pacemaker function of HCN4 channels in mice is confined to embryonic development and requires cyclic AMP. EMBO J 27: 692–703, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  33. Herrmann S, Stieber J, Ludwig A. Pathophysiology of HCN channels. Pflügers Arch 454: 517–522, 2007. [DOI] [PubMed] [Google Scholar]
  34. Horwitz GC, Lelli A, Geleoc GS, Holt JR. HCN channels are not required for mechanotransduction in sensory hair cells of the mouse inner ear. PloS One 5: e8627, 2010. [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. Keen EC, Hudspeth AJ. Transfer characteristics of the hair cell's afferent synapse. Proc Natl Acad Sci USA 103: 5537–5542, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Koch U, Grothe B. Hyperpolarization-activated current (Ih) in the inferior colliculus: distribution and contribution to temporal processing. J Neurophysiol 90: 3679–3687, 2003. [DOI] [PubMed] [Google Scholar]
  38. Kouranova EV, Strassle BW, Ring RH, Bowlby MR, Vasilyev DV. 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, 2008. [DOI] [PubMed] [Google Scholar]
  39. Kuba H, Yamada R, Fukui I, Ohmori H. Tonotopic specialization of auditory coincidence detection in nucleus laminaris of the chick. J Neurosci 25: 1924–1934, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Kubo Y, Adelman JP, Clapham DE, Jan LY, Karschin A, Kurachi Y, Lazdunski M, Nichols CG, Seino S, Vandenberg CA. International Union of Pharmacology. LIV. Nomenclature and molecular relationships of inwardly rectifying potassium channels. Pharmacol Rev 57: 509–526, 2005. [DOI] [PubMed] [Google Scholar]
  41. Lacas-Gervais S, Guo J, Strenzke N, Scarfone E, Kolpe M, Jahkel M, De Camilli P, Moser T, Rasband MN, Solimena M. BetaIVSigma1 spectrin stabilizes the nodes of Ranvier and axon initial segments. J Cell Biol 166: 983–990, 2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Leao KE, Leao RN, Sun H, Fyffe RE, Walmsley B. Hyperpolarization-activated currents are differentially expressed in mice brainstem auditory nuclei. J Physiol 576: 849–864, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Leao RN, Svahn K, Berntson A, Walmsley B. Hyperpolarization-activated (I) currents in auditory brainstem neurons of normal and congenitally deaf mice. Eur J Neurosci 22: 147–157, 2005. [DOI] [PubMed] [Google Scholar]
  44. Li GL, Keen E, Andor-Ardo D, Hudspeth AJ, von Gersdorff H. The unitary event underlying multiquantal EPSCs at a hair cell's ribbon synapse. J Neurosci 29: 7558–7568, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Llano I, Marty A, Armstrong CM, Konnerth A. Synaptic- and agonist-induced excitatory currents of Purkinje cells in rat cerebellar slices. J Physiol 434: 183–213, 1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  46. Ludwig A, Budde T, Stieber J, Moosmang S, Wahl C, Holthoff K, Langebartels A, Wotjak C, Munsch T, Zong X, Feil S, Feil R, Lancel M, Chien KR, Konnerth A, Pape HC, Biel M, Hofmann F. Absence epilepsy and sinus dysrhythmia in mice lacking the pacemaker channel HCN2. EMBO J 22: 216–224, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Magee JC. Dendritic hyperpolarization-activated currents modify the integrative properties of hippocampal CA1 pyramidal neurons. J Neurosci 18: 7613–7624, 1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Magee JC. Dendritic integration of excitatory synaptic input. Nat Rev 1: 181–190, 2000. [DOI] [PubMed] [Google Scholar]
  49. Magee JC, Johnston D. Plasticity of dendritic function. Curr Opin Neurobiol 15: 334–342, 2005. [DOI] [PubMed] [Google Scholar]
  50. Manis PB, Marx SO. Outward currents in isolated ventral cochlear nucleus neurons. J Neurosci 11: 2865–2880, 1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  51. Marcotti W, Johnson SL, Holley MC, Kros CJ. Developmental changes in the expression of potassium currents of embryonic, neonatal and mature mouse inner hair cells. J Physiol 548: 383–400, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  52. McLean WJ, Smith KA, Glowatzki E, Pyott SJ. Distribution of the Na,K-ATPase alpha subunit in the rat spiral ganglion and organ of Corti. J Assoc Res Otolaryngol 10: 37–49, 2009. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Mo Z, Davis RL. Endogenous firing patterns of murine spiral ganglion neurons. J Neurophysiol 77: 1294–1305, 1997a. [DOI] [PubMed] [Google Scholar]
  54. Mo ZL, Davis RL. Heterogeneous voltage dependence of inward rectifier currents in spiral ganglion neurons. J Neurophysiol 78: 3019–3027, 1997b. [DOI] [PubMed] [Google Scholar]
  55. Moosmang S, Stieber J, Zong X, Biel M, Hofmann F, Ludwig A. Cellular expression and functional characterization of four hyperpolarization-activated pacemaker channels in cardiac and neuronal tissues. Eur J Biochem 268: 1646–1652, 2001. [DOI] [PubMed] [Google Scholar]
  56. Moser T, Neef A, Khimich D. Mechanisms underlying the temporal precision of sound coding at the inner hair cell ribbon synapse. J Physiol 576: 55–62, 2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Nolan MF, Malleret G, Dudman JT, Buhl DL, Santoro B, Gibbs E, Vronskaya S, Buzsáki G, Siegelbaum SA, Kandel ER, Morozov A. A behavioral role for dendritic integration: HCN1 channels constrain spatial memory and plasticity at inputs to distal dendrites of CA1 pyramidal neurons. Cell 119: 719–732, 2004. [DOI] [PubMed] [Google Scholar]
  58. Nolan MF, Malleret G, Lee KH, Gibbs E, Dudman JT, Santoro B, Yin D, Thompson RF, Siegelbaum SA, Kandel ER, Morozov A. The hyperpolarization-activated HCN1 channel is important for motor learning and neuronal integration by cerebellar Purkinje cells. Cell 115: 551–564, 2003. [DOI] [PubMed] [Google Scholar]
  59. Oertel D. The role of timing in the brain stem auditory nuclei of vertebrates. Ann Rev Physiol 61: 497–519, 1999. [DOI] [PubMed] [Google Scholar]
  60. Palmer AR, Russell IJ. Phase-locking in the cochlear nerve of the guinea-pig and its relation to the receptor potential of inner hair-cells. Hear Res 24: 1–15, 1986. [DOI] [PubMed] [Google Scholar]
  61. Parks TN. The AMPA receptors of auditory neurons. Hear Res 147: 77–91, 2000. [DOI] [PubMed] [Google Scholar]
  62. Pian P, Bucchi A, Decostanzo A, Robinson RB, Siegelbaum SA. Modulation of cyclic nucleotide-regulated HCN channels by PIP(2) and receptors coupled to phospholipase C. Pflügers Arch 455: 125–145, 2007. [DOI] [PubMed] [Google Scholar]
  63. Postlethwaite M, Hennig MH, Steinert JR, Graham BP, Forsythe ID. Acceleration of AMPA receptor kinetics underlies temperature-dependent changes in synaptic strength at the rat calyx of Held. J Physiol 579: 69–84, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  64. Raman IM, Zhang S, Trussell LO. Pathway-specific variants of AMPA receptors and their contribution to neuronal signaling. J Neurosci 14: 4998–5010, 1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Rathouz M, Trussell L. Characterization of outward currents in neurons of the avian nucleus magnocellularis. J Neurophysiol 80: 2824–2835, 1998. [DOI] [PubMed] [Google Scholar]
  66. Reyes AD, Rubel EW, Spain WJ. Membrane properties underlying the firing of neurons in the avian cochlear nucleus. J Neurosci 14: 5352–5364, 1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
  67. Robertson D. Possible relation between structure and spike shapes of neurones in guinea pig cochlear ganglion. Brain Res 109: 487–496, 1976. [DOI] [PubMed] [Google Scholar]
  68. Robinson RB, Siegelbaum SA. Hyperpolarization-activated cation currents: from molecules to physiological function. Ann Rev Physiol 65: 453–480, 2003. [DOI] [PubMed] [Google Scholar]
  69. Rodrigues AR, Oertel D. Hyperpolarization-activated currents regulate excitability in stellate cells of the mammalian ventral cochlear nucleus. J Neurophysiol 95: 76–87, 2006. [DOI] [PubMed] [Google Scholar]
  70. Rothman JS, Manis PB. Differential expression of three distinct potassium currents in the ventral cochlear nucleus. J Neurophysiol 89: 3070–3082, 2003a. [DOI] [PubMed] [Google Scholar]
  71. Rothman JS, Manis PB. The roles potassium currents play in regulating the electrical activity of ventral cochlear nucleus neurons. J Neurophysiol 89: 3097–3113, 2003b. [DOI] [PubMed] [Google Scholar]
  72. Ruel J, Nouvian R, Gervais d'Aldin C, Pujol R, Eybalin M, Puel JL. Dopamine inhibition of auditory nerve activity in the adult mammalian cochlea. Eur J Neurosci 14: 977–986, 2001. [DOI] [PubMed] [Google Scholar]
  73. Santoro B, Chen S, Luthi A, Pavlidis P, Shumyatsky GP, Tibbs GR, Siegelbaum SA. Molecular and functional heterogeneity of hyperpolarization-activated pacemaker channels in the mouse CNS. J Neurosci 20: 5264–5275, 2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  74. Schneggenburger R, Forsythe ID. The calyx of Held. Cell Tissue Res 326: 311–337, 2006. [DOI] [PubMed] [Google Scholar]
  75. Scott LL, Mathews PJ, Golding NL. Posthearing developmental refinement of temporal processing in principal neurons of the medial superior olive. J Neurosci 25: 7887–7895, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  76. Seal RP, Akil O, Yi E, Weber CM, Grant L, Yoo J, Clause A, Kandler K, Noebels JL, Glowatzki E, Lustig LR, Edwards RH. Sensorineural deafness and seizures in mice lacking vesicular glutamate transporter 3. Neuron 57: 263–275, 2008. [DOI] [PMC free article] [PubMed] [Google Scholar]
  77. Shin KS, Rothberg BS, Yellen G. Blocker state dependence and trapping in hyperpolarization-activated cation channels: evidence for an intracellular activation gate. J Gen Physiol 117: 91–101, 2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
  78. Siegel JH. Spontaneous synaptic potentials from afferent terminals in the guinea pig cochlea. Hear Res 59: 85–92, 1992. [DOI] [PubMed] [Google Scholar]
  79. Siegel JH, Dallos P. Spike activity recorded from the organ of Corti. Hear Res 22: 245–248, 1986. [DOI] [PubMed] [Google Scholar]
  80. Smith AJ, Owens S, Forsythe ID. Characterisation of inhibitory and excitatory postsynaptic currents of the rat medial superior olive. J Physiol 529: 681–698, 2000. [DOI] [PMC free article] [PubMed] [Google Scholar]
  81. Stieber J, Herrmann S, Feil S, Loster J, Feil R, Biel M, Hofmann F, Ludwig A. The hyperpolarization-activated channel HCN4 is required for the generation of pacemaker action potentials in the embryonic heart. Proc Natl Acad Sci USA 100: 15235–15240, 2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  82. Szabo ZS, Harasztosi CS, Sziklai I, Szucs G, Rusznak Z. Ionic currents determining the membrane characteristics of type I spiral ganglion neurons of the guinea pig. Eur J Neurosci 16: 1887–1895, 2002. [DOI] [PubMed] [Google Scholar]
  83. Trussell LO. Synaptic mechanisms for coding timing in auditory neurons. Ann Rev Physiol 61: 477–496, 1999. [DOI] [PubMed] [Google Scholar]
  84. Tsay D, Dudman JT, Siegelbaum SA. HCN1 channels constrain synaptically evoked Ca2+ spikes in distal dendrites of CA1 pyramidal neurons. Neuron 56: 1076–1089, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. Ulens C, Tytgat J. Functional heteromerization of HCN1 and HCN2 pacemaker channels. J Biol Chem 276: 6069–6072, 2001. [DOI] [PubMed] [Google Scholar]
  86. van Welie I, Wadman WJ, van Hooft JA. Low affinity block of native and cloned hyperpolarization-activated Ih channels by Ba2+ ions. Eur J Pharmacol 507: 15–20, 2005. [DOI] [PubMed] [Google Scholar]
  87. Wainger BJ, DeGennaro M, Santoro B, Siegelbaum SA, Tibbs GR. Molecular mechanism of cAMP modulation of HCN pacemaker channels. Nature 411: 805–810, 2001. [DOI] [PubMed] [Google Scholar]
  88. Yamada R, Kuba H, Ishii TM, Ohmori H. Hyperpolarization-activated cyclic nucleotide-gated cation channels regulate auditory coincidence detection in nucleus laminaris of the chick. J Neurosci 25: 8867–8877, 2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  89. Ying SW, Jia F, Abbas SY, Hofmann F, Ludwig A, Goldstein PA. Dendritic HCN2 channels constrain glutamate-driven excitability in reticular thalamic neurons. J Neurosci 27: 8719–8732, 2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Zolles G, Klocker N, Wenzel D, Weisser-Thomas J, Fleischmann BK, Roeper J, Fakler B. Pacemaking by HCN channels requires interaction with phosphoinositides. Neuron 52: 1027–1036, 2006. [DOI] [PubMed] [Google Scholar]

Associated Data

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

Supplementary Materials

[Supplemental Figures]
00506.2009_index.html (822B, html)
00506.2009_1.pdf (268.8KB, pdf)

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

RESOURCES