Abstract
The cholinergic modulation of the electrical properties of spinal motoneurones was investigated in vitro, with the use of the whole-cell patch-clamp recording technique in lumbar spinal cord slices from juvenile urodeles (Pleurodeles waltlii). Bath application of acetylcholine (20 μm) with eserine (20 μm) induced an increase in the resting membrane potential, a decrease of the input resistance, a decrease of the action potential amplitude, and a reduction of the medium afterhyperpolarization (mAHP) that followed each action potential. Moreover, the firing rate of motoneurones during a depolarizing current pulse and the slope of their stimulus current–spike frequency relation were increased. All of these effects were mimicked by extracellular application of muscarine (20 μm), and blocked by application of the muscarinic receptor antagonist atropine (0.1–1 μm). They were not observed during bath application of nicotine (10 μm). These results suggest that the cholinergic modulation of spinal motoneurone excitability was mediated by activation of muscarinic receptors. Our results further show that the muscarinic action primarily resulted from a reduction of the Ca2+-activated K+ current responsible for the mAHP, an inhibition of the hyperpolarization-activated cation current, Ih, and an enhancement of the inward rectifying K+ current, IKir. We conclude that cholinergic modulation can contribute significantly to the production of motor behaviour by altering several ionic conductances responsible for the repetitive discharge of motoneurones.
The excitability and firing pattern of α-motoneurones are determined by voltage-gated channels, which also are targets for several neuromodulators (see Russo & Hounsgaard, 1999; Rekling et al. 2000 for reviews).
Among the neuromodulators that can contribute to the regulation of excitability of α-motoneurones, acetylcholine (ACh) is a good candidate, especially for spinal α-motoneurones. Indeed, immunohistochemical studies have shown that spinal α-motoneurones are contacted by choline acetyltransferase (ChAT)- or vesicular acetylcholine transporter (VAChT)-immunoreactive axon terminals (Fetcho, 1986; Nagy et al. 1993; Li et al. 1995; Arvidsson et al. 1997; Ichikawa & Shimizu, 1998; Schafer et al. 1998; Welton et al. 1999; Brownstone et al. 2004; Wilson et al. 2004), part of which probably originate from axon collaterals of nearby motoneurones (Cullheim et al. 1977; Perrins & Roberts, 1995). Moreover, spinal α-motoneurones in the rat express the muscarinic M2 acetylcholine receptor on their plasma membrane (Welton et al. 1999; Hellstrom et al. 2003; Wilson et al. 2004).
Despite morphological evidence of a cholinergic innervation of α-motoneurones, there have been relatively few investigations on the influences of cholinergic inputs on the excitability of spinal motoneurones. In the adult cat, iontophoretic application of ACh increases the motoneurone excitability via a reduction in a potassium conductance (Zieglgansberger & Reiter, 1974). In the neonatal rat and mouse, nicotinic and muscarinic agonists can depolarize and increase the firing of spinal motoneurones (Kurihara et al. 1993; Ogier et al. 2004). The excitatory effect of muscarinic agonists is due in part to a reduction of the Ca2+-activated K+ current (IK(Ca)) responsible for the after-hyperpolarization potential (AHP) of the action potential (Brownstone et al. 2004). In the adult turtle, muscarinic agonists increase the excitability of motoneurones, by facilitating an L-type calcium current and reducing an M-like potassium conductance (Alaburda et al. 2002).
Altogether these studies suggest that ACh modulates the excitability of spinal motoneurones, but the ionic basis of this excitation has not been fully elucidated. Understanding this mechanism is important because ACh is involved in the function of spinal locomotor networks in several vertebrate preparations (Smith et al. 1988; Panchin et al. 1991; Kiehn et al. 1996; Fok & Stein, 2002; Quinlan et al. 2004).
Among vertebrates, the juvenile urodele is a suitable experimental model for investigating the cholinergic modulation of spinal α-motoneurones. Indeed, in vitro spinal cord preparations are easy to keep alive for several hours and allow stable intracellular recordings (Wheatley & Stein, 1992; Luksch et al. 1996). Moreover, fully metamorphosed juvenile urodeles display functionally mature locomotor behaviours (Delvolvéet al. 1997; Chevallier et al. 2004a), suggesting that their α-motoneurones express functionally mature ionic currents and neuromodulation.
In the present study, we performed whole-cell recordings from identified motoneurones in slices of lumbar spinal cord of juvenile urodeles. Our results show that the cholinergic excitation of spinal motoneurones is mediated by muscarinic receptors. They further show, for the first time, that the muscarinic action primarily results from modulation of three ionic currents: the hyperpolarization-activated cation current (Ih), the IK(Ca) current and the inward rectifying K+ current (IKir).
It was concluded that through several forms of modulation, the cholinergic system is ideally placed to shape the motoneuronal discharges during motor behaviour.
A preliminary account of these results has appeared in abstract form (Chevallier et al. 2004b).
Methods
Experiments were carried out on 47 juvenile amphibian urodeles (Pleurodeles waltlii) with snout vent lengths (SVLs) ranging from 40 to 67 mm. Animals obtained from Blades Biological Ltd (UK) were kept in an aquarium at room temperature and fed twice a week. Surgical procedures, and handling and housing of the animals were in accordance with protocols approved by the INSERM Ethics Committee and conformed to NIH guidelines.
Slice preparation
Animals were deeply anaesthetized by immersion in a 0.1% aqueous solution of tricaine methanesulphonate (MS-222; Sigma). After evisceration, the spinal cord was exposed by a dorsal laminectomy in a dissection dish containing ice-cold control Ringer solution (mm: NaCl, 130; KCl, 2.1; CaCl2, 2.6; MgCl2, 0.2; Hepes, 4; glucose, 5; NaHCO3, 1), saturated with O2 (pH 7.4). A portion of the cord comprising the 14th–18th segments was isolated and longitudinal slices containing the ventral part of the cord, over two spinal segments, were prepared by carefully removing two-thirds of the most dorsal part of the cord using a microscalpel (Sharpoint; Perouse Implant). The slices obtained (around 150 μm thick) were kept in ice-cold oxygenated control Ringer solution during at least 1–2 h prior recording session.
A single slice was transferred to the thermoregulated recording chamber and viewed under infrared (IR) light using a CCD camera attached to an upright Olympus Optical fluorescent microscope BX51WI (Olympus France, Rungis, France) equipped with a 40× water immersion lens (LUMPlanFI/IR). The slice was continuously perfused at 2 ml min−1 with control oxygenated Ringer solution (see above) and maintained at 18–19°C. Ringer solution could be exchanged with drug-containing solution within 2–3 min.
Whole-cell recordings
Whole-cell recordings from IR visualized neurones were performed in either the voltage- or current-clamp configuration using 15 MΩ electrodes containing (mm): potassium gluconate, 110; KCl, 10; MgCl2, 1; EGTA–KOH, 1; Hepes, 10; CaCl2, 0.1; GTPNa3, 0.10; AMPc, 0.20; Na2-ATP, 4; and 2% biocytin. The pH was adjusted to 7.3 with KOH and the osmolarity was adjusted to 280 mosmol l−1 (i.e. 5 mosmol l−1 above the osmolarity of the perfused medium) with mannitol.
The voltage- and current-clamp recordings were made with an Axoclamp-2B intracellular amplifier (Axon Instruments, Union City, CA, USA) and monitored on an oscilloscope (DSO 630, Gould, Ilford, Essex, UK). Analog signals were filtered (0–30 kHz) and sampled (5–50 kHz) with the Digidata 1322-A analog-to-digital board (Axon Instruments) and stored on hard disk for subsequent analysis. pCLAMP 9 software (Axon Instruments) was used to control the current command outputs and to acquire data. Bridge balance and input resistance were monitored throughout the experiments by means of hyperpolarizing (−20 pA; 1.5 s) current pulses. Voltages were not corrected for the liquid junction potential which was close to 0 mV with the intracellular and various extracellular solutions used in the present experiments.
Whole-cell recordings from 68 neurones were performed in stable conditions characterized by the maintenance of a resting potential more negative than −60 mV and spike amplitude > 60 mV, thereby attesting to the absence of excessive damage to the cell.
Identification of motoneurones
All the recorded cells were intracellularly labelled after recording with biocytin in the patch pipette.
In 28 animals, a retrograde tracing technique was used in order to identify motoneurones. Two to seven days before electrophysiological experiment recording, animals were deeply anaesthetized by immersion in MS-222 as described above. A solution of carbocyanin at 15% in ethanol (Fast DiI, Molecular Probes) was injected with a Hamilton syringe into the hindlimb muscles. Then, animals were kept individually in empty tanks in which the bottom was covered with a fresh wet paper until the electrophyiological experiments. In total, 32 cells, located in the ventral horn of the lumbar spinal cord, were identified as motoneurones by their fluorescence after retrograde labelling with FDiI (Russier et al. 2002) (Fig. 1).
Figure 1. Identification of an acetylcholine-responsive spinal motoneurone.
A, photomicrograph of a longitudinal slice of lumbar spinal cord illustrating a motoneurone labelled with the retrograde dye FDiI injected in the ipsilateral hindlimb muscles, 2–7 days before recording. Lateral border of the slice on the left and medial one on the right. B, the FdiI-labelled motoneurone was injected with biocytin during the recording session. The photomicrograph in C is an overlay of the photomicrographs in A and B. Note that this motoneurone responded to ACh, as illustrated in Fig. 2A. Scale bar: 45 μm.
In 19 animals, 36 biocytin-labelled cells located in the ventral horn of the spinal cord were identified as putative motoneurones by their large soma diameter (i.e. > 20 μm) and morphology (monopolar or bipolar; Fetcho, 1986). A subset (4/36) of the biocytin-labelled neurones were reliably identified as motoneurones by the additional observation of their axon exiting the ventral root. Furthermore, 36/36 biocytin-labelled neurones displayed the characteristic-firing pattern of motoneurones (i.e. repetitive firing) in response to sustained injection of depolarizing current (Russo & Hounsgaard, 1999) and no significant difference between their intrinsic electrophysiological properties and those of FDiI backfilled motoneurones could be detected.
Biocytin labelling
After recording, longitudinal slices were incubated overnight in a fixative solution (4% paraformaldehyde in 0.1 mm phosphate buffer). Biocytin was then revealed with the following fluorescence protocol: longitudinal slices were rapidly rinsed in Tris–NaCl buffer, preincubated for 45 min in Tris–Triton X-100 and incubated for 2 h at room temperature with Alexa-fluor 488-conjugated streptavidin (Molecular Probes) diluted at 1: 200 in Tris–Triton. Spinal cord slices were finally rinsed, mounting in DAKO fluorescence mounting medium and coverslipped. Staining was observed in a Zeiss Axiophot 2 fluorescence microscope (Zeiss, Jena, Germany) equipped with the appropriate filter sets.
Solutions and drugs
In all experiments, fast excitatory and inhibitory synaptic transmission were eliminated with (±)-2-amino-5-phosphonopentanoic acid (AP-5, 50 μm; Sigma), 6-cyano-7-nitroquinoxaline-2,3-dione (CNQX, 20 μm; Tocris Cookson, Ltd, Bristol, UK) and 1 μm strychnine (Sigma). The drugs were added to the extracellular medium and continuously superfused over the preparation. Acetylcholine (ACh, 20 μm; Sigma) was freshly prepared before experiments. The cholinesterase inhibitor eserine (physostigmine, 20 μm; Sigma) was added to enhance the effect of ACh through limitation of degradation. Other drugs used were: (±)-muscarine chloride (muscarine, 20 μm; Sigma); atropine (0.1–1 μm; Sigma); nicotine (20 μm; Sigma); apamine (0.2 μm; Sigma), d-tubocurarine (10 μm; Sigma), bicuculline (20 μm; Sigma), linopirdine (20 μm; Sigma), ZD 7288 (100 μm; Tocris), caesium (1 mm) and barium (500 μm).
Data analysis
Data analysis was performed using Axograph (Axon Instruments), Microsoft Excel and SigmaPlot 8 (SPSS Inc., Chicago, IL, USA). The data were processed with the use of standard statistical analyses (SigmaStat software). Values are given as the mean ±s.e.m. and n is the number of motoneurones. Differences were considered to be significant for P < 0.05.
The passive (cell-at-rest) properties measured included resting membrane potential (in mV) and ‘steady-sate’ input resistance (in MΩ). The ‘steady-sate’ input resistance was estimated as the voltage change induced by small hyperpolarizing current pulses of 1.5 s duration applied from resting membrane potential divided by the amount of current injected.
The firing of motoneurones was characterized both by calculating the instantaneous frequency and the mean frequency during their steady state discharge evoked by a 2 s depolarizing current. For each intensity of the current pulse, the mean firing frequency was plotted against the current intensity (f–I plot). For each motoneurone, the slope of the f–I relation, which expressed the gain, was measured as the slope of the straight line fitting of the f–I plot. Unless otherwise specified, a bias current was always injected to reach the same membrane potential before and after drug application. The amplitude of the medium after-hyperpolarization (mAHP) was estimated as the difference (mV) between the threshold of the action potential and the peak of the mAHP for each action potential fired at low rate (1–2 Hz). The threshold of the action potential was visually determined from the voltage record as the membrane potential at the point of maximal change of voltage (inflection point). Muscarine-induced currents were obtained using membrane potential ramps and a subtraction procedure. I–V relations were obtained during ramps from −40 to −140 mV over 1 s. The current (Im) was averaged from four successive responses. The conductance (Gm) was determined before and after application of muscarine at −25 mV (Vm) from the measured equilibrium potential for K+ (EK) using the equation Gm=Im/(Vm–EK). EK was determined at the inflection point of the I–V relationship.
Results
Effects of acetylcholine on the excitability of spinal motoneurones
The effect of acetylcholine (ACh) was tested on 13 spinal motoneurones recorded in current-clamp mode after blockade of the fast synaptic transmission (see Methods). For each motoneurone, ACh (20 μm) was bath applied in combination with eserine (20 μm) to limit the degradation of ACh. In 9 out of the 13 recorded motoneurones, application of this medium on the spinal cord slices induced a hyperpolarization of the resting membrane potential (−6.0 ± 0.9 mV, n= 9), whereas in the remaining four cells the resting membrane potential remained unchanged.
The input resistance and the spike characteristics were monitored in a subgroup of 7 out of the 13 recorded motoneurones. ACh application induced a decrease in both the input resistance (1256 ± 183 MΩversus 1434 ± 210 MΩ, n= 7/7 cells; t test, P < 0.01) and the spike amplitude (61 ± 3 mV versus 72 ± 2 mV, n= 7/7 cells; t test, P < 0.01). By contrast, no statistically significant changes in the width of the spikes were observed (1.408 ± 0.199 s versus 1.234 ± 0.120 s, n= 7/7 cells; t test, n.s.). Unexpected from the preceding results, the excitability of 12 out of the 13 motoneurones increased (+10.7 to +80.7%), as seen by the number of spikes elicited by a depolarizing current pulse applied at the same membrane potential before and after application of ACh in combination with eserine (Fig. 2A).
Figure 2. Acetylcholine increases the excitability of motoneurones.
A, response to a depolarizing current pulse (10 pA; 2 s). The motoneurone displayed a tonic discharge both in control conditions (left) and after addition of acetylcholine (20 μm) and eserine (20 μm) to the bath (right). Furthermore, in the presence of acetylcholine and eserine the neurone exhibited an enhanced responsiveness to the depolarizing current pulse. A positive bias current was injected into the cell to adjust its membrane potential to the same value as in control conditions. B–C, plot of the instantaneous firing frequency against time before (•) and during bath application of a mixture of acetylcholine (20 μm) and eserine (20 μm) (○). In B, a current pulse (80 pA; 2 s) was injected in the neurone before and after application of drugs. In C, the same sustained discharge was induced by a pulse of 60 pA (2 s) in control condition and 40 pA (2 s) with ACh plus eserine. Arrows indicate the initial doublets of action potentials. Same neurone in A and B. In this and the following figures, the fast synaptic transmission was blocked with CNQX (20 μm), AP-5 (50 μm) and strychnine (1 μm) in the bath.
In all cells, the discharge pattern, under control or cholinergic conditions, consisted of an initial burst of action potentials at high frequency (20–60 Hz) followed by a sustained discharge with a slight adaptation (Fig. 2B). However, for similar sustained discharge, the initial burst of action potentials had a higher frequency in the presence of ACh and eserine than in control conditions (Fig. 2C).
Muscarinic receptors mediate the cholinergic increase in the excitability of spinal motoneurones
The effect of bath application of muscarine (20 μm) was tested on 22 motoneurones under current clamp conditions, in the presence of blockers of fast synaptic transmission. The excitability of 19 of these cells was increased (+3.1 to +62.0%), since the same depolarizing current pulse applied at the same potential induced more action potentials (Fig. 3A). In the remaining three cells, the excitability was not affected.
Figure 3. The excitatory effects of acetylcholine are mediated by muscarinic receptors.
A, the tonic discharge of a motoneurone to a depolarizing current pulse (10 pA; 2 s) (left) was increased during bath application of muscarine (20 μm) (middle). This effect was reversed by additional application of atropine (0.3 μm) in the medium (right). B, plot of the instantaneous firing frequency against time in control conditions (•), during bath application of muscarine (20 μm) (○) and after additional application of atropine (0.3 μm) to the bath (δ). Firing was induced by a depolarizing current pulse (90 pA; 2 s). Note that the motoneurone displayed a slight adaptation of its firing frequency in all conditions. Same neurone in A and B.
The input resistance and the spike characteristics were monitored in a fraction of 8 out of the 19 motoneurones in which the excitability was increased by muscarine. Similarly with the results obtained during bath application of ACh plus eserine, the increase in excitability induced by muscarine was associated with a decrease in the input resistance (868 ± 199 MΩversus 1011 ± 226 MΩ, n= 8/8 cells; t test, P < 0.05). Furthermore, as during ACh application, the amplitude of each action potential within a train of spikes in response to a constant-current stimulus was decreased (61 ± 3 mV versus 68 ± 2 mV, n= 8/8 cells; t test, P < 0.01). However, muscarine induced a membrane hyperpolarization (−6.7 ± 1.8 mV) in only a small fraction of motoneurones (3/19), whereas the resting membrane potential of the remaining cells (16/19) was not affected.
In 14 of 19 motoneurones, in which the excitability was increased by muscarine, the firing pattern during a step depolarization was similar under control conditions and in the presence of muscarine. In both cases, the firing rate adapted from an early maximum at the onset of depolarization to a lower steady state (‘adaptating pattern’) (Fig. 3B). By contrast, in the remaining 5/19 cells, the discharge pattern in presence of muscarine consisted of an initial adaptation at the onset of the depolarizing current pulse followed by a gradual increase in frequency reaching a maximum at the end of the pulse (‘accelerating pattern’) (not illustrated). Note that, for each firing pattern, in addition to the delayed increase in firing frequency, the initial rate of firing during the pulse was also facilitated by muscarine (Fig. 3B).
In 7/7 motoneurones, bath application of the muscarinic antagonist, atropine (0.3–1 μm), blocked the effects induced by muscarine (Fig. 3A and B). After addition of atropine to the bath, the firing frequency of motoneurones decreased progressively to control values (for example, in Fig. 3B, the mean firing frequency was 13.41 ± 0.15 Hz in control conditions and 14.54 ± 0.15 Hz in the presence of atropine). The muscarinic effect on the amplitude of action potentials was also reversed by atropine (Fig. 3A). The effects of muscarine and atropine were reversible (not illustrated).
In five motoneurones, the responses to depolarizing current pulses applied at the same membrane potential, before and after application of nicotine (10 μm), were investigated. No statistically significant (t test, P= 0.99) changes in the firing rate were observed. As an example, the mean firing rate of a typical motoneurone in response to a depolarizing pulse of 60 pA was 13.12 ± 0.19 Hz in control conditions versus 13.40 ± 0.22 Hz in the presence of nicotine. Furthermore, in the five motoneurones recorded, nicotine application did not change the resting membrane potential, the input resistance (939 ± 275 MΩversus 955 ± 258 MΩ, n= 5; t test, n.s.) or the spike amplitude (81 ± 5 mV versus 79 ± 4 mV, n= 5; paired t test, n.s.).
Taken as a whole, our results indicated that the acetylcholine effects on the excitability of motoneurones were mediated by muscarinic receptors.
Muscarine enhances the input–output relationship of motoneurones
The plot in Fig. 4A illustrates the relationship between the mean firing rate and the amplitude of depolarizing currents injected in a typical motoneurone (f–I plot). The f–I relation was linear over the whole range of stimulus intensities in control conditions (y= 0.141x+ 0.529, r2= 0.99), as well as in the presence of muscarine (y= 0.229x+1.023, r2= 0.99) (Fig. 4A) or acetylcholine and eserine (not illustrated). This was observed in motoneurones which exhibited an adapting firing pattern as well as in motoneurones which exhibited an accelerating firing pattern in the presence of muscarine. Furthermore, in most cases tested (33 of 37 cells), the slope of the f–I relationship was steeper during application of muscarine, indicating an increase in the gain (0.323 ± 0.023 Hz pA−1versus 0.266 ± 0.019 Hz pA−1; paired t test, P < 0.001). In 9 out of 9 tested motoneurones, the gain increase induced by the cholinergic agonists was reversed with atropine (0.3–1 μm; Fig. 4B).
Figure 4. Muscarine increases the gain of motoneurones.
A, frequency current plot (f–I plot) for a motoneurone. Each point represents a mean value of the steady state frequency plotted against current intensity before (•) and after (○) application of muscarine (20 μm). Muscarine shifted the f–I plot up and increased its steepness. These effects were reversed by addition of atropine (0.3 μm) to the bath (δ). In every condition, the steady state frequency varied linearly with the current intensity. B, mean value of the slope of the f–I plot for 9 motoneurones in control conditions (black bar), during bath application of muscarinic agonists (20 μm acetylcholine plus 20 μm eserine, n= 2; 20 μm muscarine, n= 7) (white bar) and after additional application of atropine (0.3–1 μm) (grey bar) to the bath. Error bars are s.e.m.***P < 0.001. n.s., not significant.
Muscarinic modulation of the medium after hyperpolarization (mAHP)
Spinal motoneurones exhibited action potentials followed by a complex AHP (Fig. 5A). The first phase of the AHP (fast AHP: fAHP) is a direct continuation of the repolarizing phase of the action potential. The second phase peaked 15–50 ms after the action potential and displayed a more prolonged time course (20–150 ms). This component was referred to as the medium AHP (mAHP) (Barrett et al. 1980). The two phases were separated by a depolarizing waveform, the afterdepolarization (ADP). Because the mAHP was the most frequently observed and the fAHP did not always appear, we focused on the cholinergic modulation of the mAHP. As described in other species (Goh & Pennefather, 1987; Zhang & Krnjevic, 1987; Hounsgaard et al. 1988; Seutin et al. 1997) and depicted in Fig. 5B, the mAHP of salamander motoneurones was strongly reduced (−32.3 ± 2.3%, n= 3) during application of an SK blocker (i.e. apamine, d-tubocurarine or quaternary salt of bicuculline) (Fig. 5B).
Figure 5. Muscarine decreases the amplitude of the medium after-hyperpolarization (mAHP).
A, spinal motoneurones express 3 types of afterpotentials: fast afterhyperpolarization (fAHP), afterdepolarization (ADP) and medium afterhyperpolarization (mAHP). B, superimposition of two action potentials fired by the same neurone in control conditions, and during application of apamine (0.4 μm). C, superimposition of three action potentials fired by the same neurone in control conditions, during application of muscarine (20 μm) and after additional application of atropine (0.3 μm) to the bath. Each action potential was elicited by a current pulse of low intensity (10 pA). D, bar graph representing the mean mAHP amplitude (n= 9 neurones) in control conditions (black bar), during bath application of muscarine (20 μm) (white bar) and after additional application of atropine (0.3–1 μm) (grey bar). mAHP amplitude was determined between action potential threshold and peak of the mAHP (dotted lines in C). ***P < 0.001. n.s., not significant.
Bath application of ACh plus eserine (n= 5 cells) or muscarine (n= 8 cells) clearly decreased the mAHP amplitude, as exemplified in Fig. 5C. On average, cholinergic agonists decreased significantly (t test, P < 0.001) the mAHP from 20.8 ± 0.8 mV in control conditions to 15.4 ± 0.9 mV in the presence of cholinergic agonists (n= 9; Fig. 5D). This effect was fully reversed by atropine (20.1 ± 1.1 mV, n= 9; Fig. 5D).
Previous reports have shown that a voltage-activated K+ current, the M-current, can contribute to the mAHP (Storm, 1989), to the discharge adaptation (Wang & McKinnon, 1995) and to the control of the motoneurone gain (Alaburda et al. 2002). Because muscarinic receptors inhibited the M-current in several type of neurones (Adams et al. 1982; Halliwell & Adams, 1982), we tested the effects of linopirdine (20 μm), a specific blocker of the M-current (Aiken et al. 1995; Schnee & Brown, 1998). In all tested motoneurones (n= 4), we did not observe any change in the resting membrane potential and input resistance (594 ± 61 MΩversus 543 ± 48 MΩ, n= 4; t test n.s.). Also the mean firing frequency during a given current pulse applied at the same membrane potential was not significantly affected (t test, P= 0.91). As an example, the firing rate in a motoneurone was 25.69 ± 0.46 Hz in the presence of linopirdine versus 25.54 ± 0.37 Hz in control conditions, for a 140 pA depolarizing pulse applied at −60 mV. These results indicated that the M-current could not account for the muscarinic effects observed in the present experiments.
Altogether, our findings suggested that the increase in firing frequency and gain of motoneurone could partly be explained by a muscarinic inhibition of the mAHP, via a negative modulation of the SK channels.
Negative modulation of Ih
In response to hyperpolarizing current injection, 27 out of 28 motoneurones tested expressed a pronounced depolarizing sag during the pulse of current followed by a rebound depolarization at the termination of the stimulus (Fig. 6A and C). Depending on the amplitude of the hyperpolarizing current, the post-inhibitory rebound (PIR), either was subthreshold (Fig. 6A and C) or triggered an action potential (not illustrated). Furthermore, as exemplified in Fig. 6, the amplitude of the depolarizing sag (filled symbols in Fig. 6B) and that of the PIR (black bars in Fig. 6D) increased when the amplitude of the hyperpolarizing pulse current increased.
Figure 6. The motoneurones express Ih.
A, response of a motoneurone to a hyperpolarizing current pulse (−80 pA; 1.5 s) in the absence (control) and in the presence of Cs+ (1 mm). The peak amplitude of the hyperpolarizing voltage responses (• and ○) and the steady-state voltage amplitude reached just before the offset of the hyperpolarizing current pulse (▾ and ▿) were measured and plotted against the intensity of the hyperpolarizing current pulse (B). C, response of a motoneurone to a hyperpolarizing current pulse (−60 pA; 1.5 s) in the absence and in the presence of ZD7288 (100 μm). D, plot of the amplitude of the post-inhibitory rebound (PIR) before (black bars) and after (white bars) application of ZD7288 (100 μm) against intensity of the hyperpolarizing current pulse. In the presence of Cs+ or ZD 7288, the resting membrane potential was set to the same value (−60 mV) as in control conditions using a positive bias current. Same cell in A and B; same cell in C and D.
In all motoneurones tested, bath application of caesium (Cs+) at low concentration (1 mm) abolished or strongly reduced (−95 ± 2%, n= 6) the depolarizing sag (Fig. 6B, open symbols) and the PIR (−96 ± 3%, n= 6). At the same time, the resting membrane potential was hyperpolarized (−15 ± 1 mV, n= 7) and the input resistance was increased (1128 ± 218 MΩversus 778 ± 147 MΩ, n= 6; t test, P < 0.001). This sensitivity to low concentrations of Cs+ suggested the presence of an Ih current (Takahashi, 1990; Pape, 1996). Consistently, the PIR was abolished during the application of a low concentration of Cs+, as shown in the example in Fig. 6A. In order to further characterize the current involved in the depolarizing sag and PIR, we applied to the bath ZD 7288 (100 μm, n= 7), a selective and irreversible pharmacological blocker of Ih (BoSmith et al. 1993; Harris & Constanti, 1995; Williams et al. 1997; Hughes et al. 1998; Khakh & Henderson, 1998; Williams et al. 2002). In all cells tested, bath application of ZD 7288 strongly reduced the depolarizing sag (−81 ± 9%, n= 5) and the PIR (−88 ± 5%, n= 5; Fig. 6D). Moreover, ZD 7288, like low concentrations of Cs+, induced a membrane hyperpolarization (−9 ± 1 mV, n= 5) and a significant increase in the input resistance (1329 ± 166 MΩversus 928 ± 71 MΩ, n= 4; t test, P < 0.001). Altogether these results indicated that the spinal motoneurones of salamander expressed an Ih current.
Figure 7A illustrates a typical example of the responses to a square pulse of hyperpolarizing current recorded before (left) and after (middle) addition of muscarine (20 μm). During application of muscarine, both the peak voltage and the PIR of the response decreased. Figure 7A (right) further shows that in the presence of atropine (0.3 μm), the effects of muscarine were partly reversed.
Figure 7. Modulation of the sag and post-inhibitory rebound by muscarine.
A, response of a motoneurone to a hyperpolarizing current pulse in control conditions (left), in the presence of muscarine (20 μm) (middle) and during an additional application of atropine (0.3 μm) in the medium (right). B, plot of the mean amplitude of the sag against the peak amplitude of the hyperpolarizing voltage response. Black bars: control conditions; white bars: in the presence of muscarine (20 μm). C, plot of the amplitude of the mean post-inhibitory rebound (PIR) against the peak amplitude of the hyperpolarizing voltage response. Black bars: control conditions; white bars: in the presence of muscarine. The inset on the right shows that amplitudes of sag and PIR were measured for the same peak potential reached in control conditions and in the presence of cholinergic agonists. In B and C, the numbers above the bars indicate the numbers of cells. Error bars are s.e.m.t test, *P < 0.05, **P < 0.01.
Since the depolarizing sag and the PIR are voltage dependent, both increased, as the peak voltage reached during a hyperpolarizing pulse was more negative (black bars in Fig. 7B and C). Therefore, in order to determine the muscarinic effects, we compared the amplitude of the depolarizing sag and of the PIR, before and after application of muscarine, while adjusting the current amplitude to reach the same peak potential in the different pharmacological conditions (inset in Fig. 7). The bar graphs in Fig. 7B and C show that bath application of muscarine and ACh plus eserine decreased significantly the depolarizing sag and the PIR, respectively. These results indicated that muscarine inhibited the Ih current in spinal motoneurones.
Positive modulation of IKir
As already mentioned, activation of muscarinic receptors in salamander motoneurones caused membrane hyperpolarization and decrease in input resistance, suggesting the activation of a K+ current. A possible target for the cholinergic modulation could be an inward rectifier potassium current (IKir) (Sakmann et al. 1983; Surprenant & North, 1988). To test the hypothesis that muscarine activated a Kir conductance in motoneurones, we voltage-clamped the cells and recorded current responses to hyperpolarizing voltage ramps in the absence and in the presence of muscarine. In control conditions, the voltage ramp from −40 to −140 mV induced an inward rectifying current characterized by a steeper slope of the I–V plot at a potential more negative than −100 ± 3 mV (n= 17) (Fig. 8A), i.e. at a potential close to the K+ equilibrium potential (EK predicted to be −103 mV with the intra- and extracellular solutions used) as expected for a Kir conductance. Addition of muscarine (20 μm) to the bath enhanced the inward rectification (Fig. 8A, compare black and red traces). Substraction of the two current traces yielded the I–V relationship of the muscarine-induced current (Fig. 8B), which displayed strong inward rectification and a reversal potential (around −95 mV) near the calculated EK. The I–V relationship of the muscarine-induced current in Fig. 8B shows that muscarine enhanced the inward rectification by increasing the inward current, while exerting a relatively weak effect on the outward current. To quantify the action of muscarine on the inward rectifying current, we calculated the conductance before and after application of the drug at −25 mV with respect to the inflection point in eight motoneurones. The conductance in the presence of muscarine was significantly increased compared to the control condition (17.03 ± 4.55 nS versus 14.48 ± 3.74 nS, n= 8; Wilcoxon test, P < 0.01) (Fig. 8C). The conductance increase in the presence of muscarine was 16.73 ± 3.64% (χ2 test, n= 8, P < 0.001).
Figure 8. Modulation of a Kir current.
A, I–V relations obtained during voltage ramps in control condition (black trace), in the presence of 20 μm muscarine in the bath (red trace) and after additional bath application of 500 μm external barium (green trace). B, the muscarine current was isolated by subtracting the control current from the current recorded in the presence of muscarine. C, conductance increase (as a percentage of control) in the presence of muscarine (white bar) compared to control condition (black bar) measured at −25 mV from the measured equilibrium potential for K+. ***P < 0.001, χ2-test, n= 8.
To further characterize the involvement of a Kir conductance we tested for the sensitivity to a low concentration of barium (Ba2+). The inward current induced by muscarine was blocked at a bath concentration of 500 μm Ba2+ (Fig. 8A, green trace) (n= 6). This high sensitivity to Ba2+ is characteristic of IKir (Hagiwara et al. 1978; Sodickson & Bean, 1996). Altogether, these results indicated that activation of muscarinic receptors exerted a positive modulation of IKir.
Discussion
In the present study, we have shown a muscarinic modulation of the excitability of spinal motoneurones in the juvenile salamander. A muscarinic regulation of the spinal motoneurone excitability has previously been reported in the adult turtle (Alaburda et al. 2002; Hornby et al. 2002a, b), in the neonate rat (Kurihara et al. 1993) and in the newborn mouse (Brownstone et al. 2004).
Previous immunohistochemical studies in the adult rat have shown that spinal motoneurones express the muscarinic receptor type 2 (M2) on their somata and proximal dendrites (Welton et al. 1999; Hellstrom et al. 2003). This suggests that part of the muscarinic effects on adult spinal motoneurones is exerted through the action of M2 receptors. Recent electrophysiological data obtained in the juvenile mouse support this view (Brownstone et al. 2004). However, the subtypes of muscarinic receptors increasing the excitability of salamander motoneurones remain to be identified.
The presence of functional nicotinic receptors has been reported in spinal motoneurones in immature rats (Blake et al. 1987; Kurihara et al. 1993; Ogier et al. 2004) and in presumed motoneurones in the Xenopus laevis embryo spinal cord (Perrins & Roberts, 1994). Our data did not evidence nicotinic modulation of the excitability of spinal motoneurones in the juvenile salamander. This supports the view that the expression of nicotinic receptors in spinal motoneurones is developmentally regulated (Keiger et al. 2003; Ogier et al. 2004).
Our results further show that the activation of muscarinic receptors inhibits a Ca2+-activated K+ current (IK(Ca)) and the hyperpolarization-activated cation current (Ih), while it enhances the inward rectifying K+ current (IKir).
It has been reported that muscarine can increase the excitability of turtle spinal motoneurones by facilitating an L-type calcium current and also by decreasing the maximal conductance of an M-type potassium current (Alaburda et al. 2002). The present study ruled out the involvement of an M-type potassium current (IM) in the muscarinic regulation of excitability in spinal motoneurone in salamander. The involvement of an L-type calcium current remains to be studied, however.
Muscarinic modulation of the mAHP
Our results indicate that the mAHP that follows each action potential was mediated by Ca2+-activated K+ channels of the SK type since it was sensitive to apamine, d-tubocurarine and quaternary salts of bicuculline (for review see Vogalis et al. 2003). However, we cannot exclude the involvement of other K+ channels (Sah, 1996; Vogalis et al. 2003). Indeed, in lamprey spinal motoneurones, 20% of the sAHP (corresponding to the mAHP in the present study) are mediated by calcium-insensitive channels that might be Na+-activated K+ channels (Cangiano et al. 2002). Futhermore, in rat hippocampal neurones, several ionic currents participate in the mAHP generation (Storm, 1989). One of them, IM, may be the main target for the blocking action by muscarinic agonists on the mAHP (Halliwell & Adams, 1982). However, since we did not evidence IM during our experiments, this current does not contribute to the mAHP in spinal motoneurones of salamander.
We have evidenced a negative regulation of the mAHP by muscarinic agonists, like in rat hypoglossal and mouse spinal motoneurones (Lape & Nistri, 2000; Brownstone et al. 2004). This effect could result from either a direct negative modulation of the SK channels or from a negative modulation of calcium channels which in turn reduce the amplitude of the mAHP. Previous studies have reported such an indirect mechanism for the sAHP in lamprey spinal neurones (Cangiano et al. 2002) or in caudal Raphe neurones (Bayliss et al. 1994). In the present study, however, we did not observe variations in action potential duration in the presence of muscarinic agonists, thus indicating that muscarinic receptors probably modulated the SK channels directly.
Muscarinic modulation of repetitive firing
Several studies have shown the role of SK channels and of the IM current in controlling spike frequency adaptation in Vertebrate motoneurones (Russo & Hounsgaard, 1999; Alaburda et al. 2002). Since in the present study IM current was not detected, it is likely that firing adaptation was mainly controlled by SK channels. Furthermore, the decrease in the AHP amplitude observed in the present study probably accounted for the observed muscarine-induced increase in firing frequency.
In a few salamander motoneurones (5/22), muscarine promoted an acceleration, rather than an adaptation, of their discharge with constant-current stimulus. This behaviour, which is also present in turtle motoneurones (Svirskis & Hounsgaard, 1997; Hornby et al. 2002a), has been related to the dynamic properties underlying the generation of plateau potentials (Russo & Hounsgaard, 1999) for review). More experiments are necessary to further substantiate this idea.
In vitro experiments showed that suppression of the AHP by apamine or reduction of the AHP amplitude by 5-HT (Hounsgaard & Kiehn, 1989) induced an increase in the gain of motoneurones. A recent dynamic clamp study on cat lumbar motoneurones has demonstrated that the gain of motoneurones is inversely proportional to the AHP conductance and to the AHP time constant (Manuel et al. 2004). Therefore, it is likely that the muscarinic control of the input–output relationship in spinal motoneurones in salamander is mediated by modulation of the mAHP.
The doublet of action potentials recorded at the beginning of the response to a depolarizing current pulse reached a higher frequency in the presence of cholinergic agonists. This effect can contribute to the production of a marked and persistent enhancement of the force output. Indeed, insertion of a short, high-frequency burst into an otherwise low-frequency stimulus train can produce remarkably long-lasting enhancement of isometric force output in non-twitch muscles of arthropods (Blaschko et al. 1931; Wilson & Larimer, 1968). This ‘catch property’ was also evidenced in mammalian motor units where the output force is modulated primarily by the pattern and not by the average frequency of the input train (Burke, 1981).
Our data further evidenced a decrease in amplitude of the spikes during the activation of the muscarinic receptors. One possible explanation is that muscarine decreased the spike amplitude simply because it induced a faster firing frequency. This faster frequency resulted in a more depolarized membrane potential between spikes (see, e.g. Fig. 3A). As a result, there was probably less relief from Na+ channel inactivation, thereby decreasing spike height. However, if this explanation is correct, the first spike in the train should have a greater height than the subsequent spikes. Since this was not the case, a direct muscarinic effect on Na+ channels is probably involved.
Muscarinic modulation of inward rectifying currents
Our observation that application of Cs+ at low concentration or of ZD 7288 induced a hyperpolarization of the membrane potential suggests that Ih was tonically activated at rest and that it therefore contributed to the resting membrane potential in salamander motoneurones. The contribution of Ih to the resting membrane potential of spinal motoneurones has previously been evidenced in the newborn rat (Kjaerulff & Kiehn, 2001).
In addition to its inhibitory action on Ih, we have shown that activation of muscarinic receptors enhanced inward rectification of the I–V plot in motoneurones, via the activation of Kir channels. A likely candidate is the G-protein-activated inwardly rectifying K+ channel (Kir3) (Pfaffinger et al. 1985; Dascal, 1997). Spinal motoneurones from newborn rats (Takahashi, 1990; Kjaerulff & Kiehn, 2001) possess IKir too.
In control conditions, motoneurones displayed an inward rectifying current at hyperpolarizing potentials around EK, indicating the expression of an inward rectifying K+ current, IKir, in salamander motoneurones. This raises the possibility that Kir3 channels displayed some degree of spontaneous background activity. This background activity could be due to a constitutive Kir3 channel activity in the absence of G protein activation (Takigawa & Alzheimer, 1999; Chen & Johnston, 2005). However, we cannot exclude the presence of ambient modulators of Kir3 channels such as ACh, GABA (Derjean et al. 2003) or adenosine, which could play a role in keeping these channels spontaneously opened (Takigawa & Alzheimer, 2002). On the other hand, the salamander motoneurones could express the Ba2+-sensitive Kir2 channels (Topert et al. 1998), which are constitutively active and contribute to the resting K+ conductance in many cells (Isomoto et al. 1997). However, the fact that in our study the bath application of Ba2+ did not change the resting potential of the cell rules out this possibility.
It is well known that activation of M2 muscarinic receptors leads to an increase of a Kir3 conductance in mammal atrial myocytes (Sakmann et al. 1983; Surprenant & North, 1988) via Gβγ subunits (Pfaffinger et al. 1985). In the central nervous system, evidence for a similar muscarinic action via M2 receptors has been reported from thalamic reticular neurones (McCormick & Prince, 1986) and from interneurones of striatum (Calabresi et al. 1998), neocortex (Xiang et al. 1998) and hippocampus (McQuiston & Madison, 1999). In the present study, the type of muscarinic receptors involved in the enhancement of the Kir conductance remains to be determined.
A dynamic clamp study in newborn rats suggests that the main action of Ih during an NMDA-induced locomotor activity is to increase the firing probability in motoneurones via a tonic depolarization (Kiehn et al. 2000). Several lines of evidence suggest an involvement of ACh in the modulation of ongoing locomotor activity in vertebrates (Panchin et al. 1991; Fok & Stein, 2002; Quinlan et al. 2004). Therefore, the negative regulation of Ih by the activation of muscarinic receptors evidenced in the present study will lead to a decrease in the firing probability of spinal motoneurone excitability during ongoing locomotor activity.
Our study reveals a mechanism that may amplify and prolong hyperpolarizing events. Because the activation and deactivation kinetics of Kir3 current are much faster than those of Ih, muscarine-induced Kir3 current can rapidly hyperpolarize the neurone in response to hyperpolarizing events. The inhibition of Ih by muscarine can then delay the recovery from hyperpolarizations and thus prolonged their duration. The relatively weak facilitation of IKir by muscarine in the range of potentials more positive than EK suggests a minor role of muscarine modulation of IKir on motoneurone firing. Interestingly, a similar conclusion has previously been drawn for 5-HT modulation of IKir in neonatal rat motoneurones (Kjaerulff & Kiehn, 2001).
In conclusion, the present study reveals that muscarine induces two opposite effects on the excitability of salamander motoneurones. Indeed, the motoneuronal excitability could be increased by an inhibition of the mAHP while it could be decreased by an inhibition of Ih and an enhancement of IKir. Because these two effects are voltage dependent, muscarinic modulation can strengthen the contrast of synaptic effects by minimizing low activations and enhancing strong activations (Seeger & Alzheimer, 2001).
Acknowledgments
This work was supported in part by the ACI ‘Integrated and Computational Neuroscience’ of the Ministère déléguéà la Recherche et aux Nouvelles Technologies. S.C. received a studentship from the Ministère de l'Education Nationale, de la Recherche et de la Technologie and from Fondation pour la Recherche Médicale.
References
- Adams PR, Brown DA, Constanti A. M-currents and other potassium currents in bullfrog sympathetic neurones. J Physiol. 1982;330:537–572. doi: 10.1113/jphysiol.1982.sp014357. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aiken SP, Lampe BJ, Murphy PA, Brown BS. Reduction of spike frequency adaptation and blockade of M-current in rat CA1 pyramidal neurones by linopirdine (DuP 996), a neurotransmitter release enhancer. Br J Pharmacol. 1995;115:1163–1168. doi: 10.1111/j.1476-5381.1995.tb15019.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Alaburda A, Perrier JF, Hounsgaard J. An M-like outward current regulates the excitability of spinal motoneurones in the adult turtle. J Physiol. 2002;540:875–881. doi: 10.1113/jphysiol.2001.015982. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arvidsson U, Riedl M, Elde R, Meister B. Vesicular acetylcholine transporter (VAChT) protein: a novel and unique marker for cholinergic neurons in the central and peripheral nervous systems. J Comp Neurol. 1997;378:454–467. [PubMed] [Google Scholar]
- Barrett EF, Barrett JN, Crill WE. Voltage-sensitive outward currents in cat motoneurones. J Physiol. 1980;304:251–276. doi: 10.1113/jphysiol.1980.sp013323. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bayliss DA, Viana F, Bellingham MC, Berger AJ. Characteristics and postnatal development of a hyperpolarization-activated inward current in rat hypoglossal motoneurons in vitro. J Neurophysiol. 1994;71:119–128. doi: 10.1152/jn.1994.71.1.119. [DOI] [PubMed] [Google Scholar]
- Blake JF, Evans RH, Smith DA. The effect of nicotine on motoneurones of the immature rat spinal cord in vitro. Br J Pharmacol. 1987;90:167–173. doi: 10.1111/j.1476-5381.1987.tb16837.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blaschko HM, Castell M, Kahn JL. On the nature of the two types of response in the neuromuscular system of the crustacean amphibian claw. J Physiol. 1931;73:25–35. doi: 10.1113/jphysiol.1931.sp002793. [DOI] [PMC free article] [PubMed] [Google Scholar]
- BoSmith RE, Briggs I, Sturgess NC. Inhibitory actions of ZENECA ZD7288 on whole-cell hyperpolarization activated inward current (If) in guinea-pig dissociated sinoatrial node cells. Br J Pharmacol. 1993;110:343–349. doi: 10.1111/j.1476-5381.1993.tb13815.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brownstone RM, Hartley R, Todd AJ, Miles GB. A spinal cholinergic mechanism mediating increased motoneuron excitability. Soc Neurosci Abs 2004. 2004:875.17. [Google Scholar]
- Burke RE. Motor units: anatomy, physiology, and functional organisation. In: Brookhart JM, Mountcastle VB, Brooks VB, Geiger SR, editors. Handbook of Physiology, section 1, The Nervous System, vol. II, Motor Control. Bethesda, MD, USA: American physiological Society; 1981. pp. 345–422. [Google Scholar]
- Calabresi P, Centonze D, Pisani A, Sancesario G, North RA, Bernardi G. Muscarinic IPSPs in rat striatal cholinergic interneurones. J Physiol. 1998;510:421–427. doi: 10.1111/j.1469-7793.1998.421bk.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cangiano L, Wallen P, Grillner S. Role of apamin-sensitive KCa channels for reticulospinal synaptic transmission to motoneuron and for the afterhyperpolarization. J Neurophysiol. 2002;88:289–299. doi: 10.1152/jn.2002.88.1.289. [DOI] [PubMed] [Google Scholar]
- Chen X, Johnston D. Constitutively active G-protein-gated inwardly rectifying K+ channels in dendrites of hippocampal CA1 pyramidal neurons. J Neurosci. 2005;25:3787–3792. doi: 10.1523/JNEUROSCI.5312-04.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chevallier S, Landry M, Nagy F, Cabelguen JM. Recovery of bimodal locomotion in the spinal-transected salamander, Pleurodeles waltlii. Eur J Neurosci. 2004a;20:1995–2007. doi: 10.1111/j.1460-9568.2004.03671.x. [DOI] [PubMed] [Google Scholar]
- Chevallier S, Nagy F, Cabelguen J-M. Cholinergic control of motoneurones excitability in Salamander. Soc Neurosci Abs 2004. 2004b:410. [Google Scholar]
- Cullheim S, Kellerth JO, Conradi S. Evidence for direct synaptic interconnections between cat spinal alpha-motoneurons via the recurrent axon collaterals: a morphological study using intracellular injection of horseradish peroxidase. Brain Res. 1977;132:1–10. doi: 10.1016/0006-8993(77)90702-8. [DOI] [PubMed] [Google Scholar]
- Dascal N. Signalling via the G protein-activated K+ channels. Cell Signal. 1997;9:551–573. doi: 10.1016/s0898-6568(97)00095-8. [DOI] [PubMed] [Google Scholar]
- Delvolvé I, Bem T, Cabelguen JM. Epaxial and limb muscle activity during swimming and terrestrial stepping in the adult newt, Pleurodeles waltl. J Neurophysiol. 1997;78:638–650. doi: 10.1152/jn.1997.78.2.638. [DOI] [PubMed] [Google Scholar]
- Derjean D, Bertrand S, Le Masson G, Landry M, Morisset V, Nagy F. Dynamic balance of metabotropic inputs causes dorsal horn neurons to switch functional states. Nat Neurosci. 2003;6:274–281. doi: 10.1038/nn1016. [DOI] [PubMed] [Google Scholar]
- Fetcho JR. The organization of the motoneurons innervating the axial musculature of vertebrates. I. Goldfish (Carassius auratus) and mudpuppies (Necturus maculosus) J Comp Neurol. 1986;249:521–550. doi: 10.1002/cne.902490408. [DOI] [PubMed] [Google Scholar]
- Fok M, Stein RB. Effects of cholinergic and noradrenergic agents on locomotion in the mudpuppy (Necturus maculatus) Exp Brain Res. 2002;145:498–504. doi: 10.1007/s00221-002-1125-6. [DOI] [PubMed] [Google Scholar]
- Goh JW, Pennefather PS. Pharmacological and physiological properties of the after-hyperpolarization current of bullfrog ganglion neurones. J Physiol. 1987;394:315–330. doi: 10.1113/jphysiol.1987.sp016872. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hagiwara S, Miyazaki S, Moody W, Patlak J. Blocking effects of barium and hydrogen ions on the potassium current during anomalous rectification in the starfish egg. J Physiol. 1978;279:167–185. doi: 10.1113/jphysiol.1978.sp012338. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Halliwell JV, Adams PR. Voltage-clamp analysis of muscarinic excitation in hippocampal neurons. Brain Res. 1982;250:71–92. doi: 10.1016/0006-8993(82)90954-4. [DOI] [PubMed] [Google Scholar]
- Harris NC, Constanti A. Mechanism of block by ZD 7288 of the hyperpolarization-activated inward rectifying current in guinea pig substantia nigra neurons in vitro. J Neurophysiol. 1995;74:2366–2378. doi: 10.1152/jn.1995.74.6.2366. [DOI] [PubMed] [Google Scholar]
- Hellstrom J, Oliveira AL, Meister B, Cullheim S. Large cholinergic nerve terminals on subsets of motoneurons and their relation to muscarinic receptor type 2. J Comp Neurol. 2003;460:476–486. doi: 10.1002/cne.10648. [DOI] [PubMed] [Google Scholar]
- Hornby TG, McDonagh JC, Reinking RM, Stuart DG. Effects of excitatory modulation on intrinsic properties of turtle motoneurons. J Neurophysiol. 2002a;88:86–97. doi: 10.1152/jn.2002.88.1.86. [DOI] [PubMed] [Google Scholar]
- Hornby TG, McDonagh JC, Reinking RM, Stuart DG. Motoneurons: a preferred firing range across vertebrate species. Muscle Nerve. 2002b;25:632–648. doi: 10.1002/mus.10105. [DOI] [PubMed] [Google Scholar]
- Hounsgaard J, Kiehn O. Serotonin-induced bistability of turtle motoneurones caused by a nifedipine-sensitive calcium plateau potential. J Physiol. 1989;414:265–282. doi: 10.1113/jphysiol.1989.sp017687. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hounsgaard J, Kiehn O, Mintz I. Response properties of motoneurones in a slice preparation of the turtle spinal cord. J Physiol. 1988;398:575–589. doi: 10.1113/jphysiol.1988.sp017058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hughes SW, Cope DW, Crunelli V. Dynamic clamp study of Ih modulation of burst firing and delta oscillations in thalamocortical neurons in vitro. Neuroscience. 1998;87:541–550. doi: 10.1016/s0306-4522(98)00170-5. [DOI] [PubMed] [Google Scholar]
- Ichikawa T, Shimizu T. Organization of choline acetyltransferase-containing structures in the cranial nerve motor nuclei and spinal cord of the monkey. Brain Res. 1998;779:96–103. doi: 10.1016/s0006-8993(97)01090-1. [DOI] [PubMed] [Google Scholar]
- Isomoto S, Kondo C, Kurachi Y. Inwardly rectifying potassium channels: their molecular heterogeneity and function. Jpn J Physiol. 1997;47:11–39. doi: 10.2170/jjphysiol.47.11. [DOI] [PubMed] [Google Scholar]
- Keiger CJ, Prevette D, Conroy WG, Oppenheim RW. Developmental expression of nicotinic receptors in the chick and human spinal cord. J Comp Neurol. 2003;455:86–99. doi: 10.1002/cne.10468. [DOI] [PubMed] [Google Scholar]
- Khakh BS, Henderson G. Hyperpolarization-activated cationic currents (Ih) in neurones of the trigeminal mesencephalic nucleus of the rat. J Physiol. 1998;510:695–704. doi: 10.1111/j.1469-7793.1998.00695.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kiehn O, Johnson BR, Raastad M. Plateau properties in mammalian spinal interneurons during transmitter-induced locomotor activity. Neuroscience. 1996;75:263–273. doi: 10.1016/0306-4522(96)00250-3. [DOI] [PubMed] [Google Scholar]
- Kiehn O, Kjaerulff O, Tresch MC, Harris-Warrick RM. Contributions of intrinsic motor neuron properties to the production of rhythmic motor output in the mammalian spinal cord. Brain Res Bull. 2000;53:649–659. doi: 10.1016/s0361-9230(00)00398-1. [DOI] [PubMed] [Google Scholar]
- Kjaerulff O, Kiehn O. 5-HT modulation of multiple inward rectifiers in motoneurons in intact preparations of the neonatal rat spinal cord. J Neurophysiol. 2001;85:580–593. doi: 10.1152/jn.2001.85.2.580. [DOI] [PubMed] [Google Scholar]
- Kurihara T, Suzuki H, Yanagisawa M, Yoshioka K. Muscarinic excitatory and inhibitory mechanisms involved in afferent fibre-evoked depolarization of motoneurones in the neonatal rat spinal cord. Br J Pharmacol. 1993;110:61–70. doi: 10.1111/j.1476-5381.1993.tb13772.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lape R, Nistri A. Current and voltage clamp studies of the spike medium afterhyperpolarization of hypoglossal motoneurons in a rat brain stem slice preparation. J Neurophysiol. 2000;83:2987–2995. doi: 10.1152/jn.2000.83.5.2987. [DOI] [PubMed] [Google Scholar]
- Li W, Ochalski PA, Brimijoin S, Jordan LM, Nagy JI. C-terminals on motoneurons: electron microscope localization of cholinergic markers in adult rats and antibody-induced depletion in neonates. Neuroscience. 1995;65:879–891. doi: 10.1016/0306-4522(94)00511-3. [DOI] [PubMed] [Google Scholar]
- Luksch H, Walkowiak W, Munoz A, Ten Donkelaar HJ. The use of in vitro preparations of the isolated amphibian central nervous system in neuroanatomy and electrophysiology. J Neurosci Meth. 1996;70:91–102. doi: 10.1016/S0165-0270(96)00107-0. [DOI] [PubMed] [Google Scholar]
- McCormick DA, Prince DA. Acetylcholine induces burst firing in thalamic reticular neurones by activating a potassium conductance. Nature. 1986;319:402–405. doi: 10.1038/319402a0. [DOI] [PubMed] [Google Scholar]
- McQuiston AR, Madison DV. Muscarinic receptor activity has multiple effects on the resting membrane potentials of CA1 hippocampal interneurons. J Neurosci. 1999;19:5693–5702. doi: 10.1523/JNEUROSCI.19-14-05693.1999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Manuel M, Zytnichi D, Meunier C, Donnet M. How the AHP current controls the discharge of cat lumbar motoneurons: a study in dynamic clamp. FENS Abs. 2004;2:A058.11. [Google Scholar]
- Nagy JI, Yamamoto T, Jordan LM. Evidence for the cholinergic nature of C-terminals associated with subsurface cisterns in alpha-motoneurons of rat. Synapse. 1993;15:17–32. doi: 10.1002/syn.890150103. [DOI] [PubMed] [Google Scholar]
- Ogier R, Liu X, Tribollet E, Bertrand D, Raggenbass M. Identified spinal motoneurons of young rats possess nicotinic acetylcholine receptors of the heteromeric family. Eur J Neurosci. 2004;20:2591–2597. doi: 10.1111/j.1460-9568.2004.03746.x. [DOI] [PubMed] [Google Scholar]
- Panchin Y, Perrins RJ, Roberts A. The action of acetylcholine on the locomotor central pattern generator for swimming in Xenopus embryos. J Exp Biol. 1991;161:527–531. doi: 10.1242/jeb.161.1.527. [DOI] [PubMed] [Google Scholar]
- Pape HC. Queer current and pacemaker: the hyperpolarization-activated cation current in neurons. Annu Rev Physiol. 1996;58:299–327. doi: 10.1146/annurev.ph.58.030196.001503. [DOI] [PubMed] [Google Scholar]
- Perrins R, Roberts A. Nicotinic and muscarinic ACh receptors in rhythmically active spinal neurones in the Xenopus laevis embryo. J Physiol. 1994;478:221–228. doi: 10.1113/jphysiol.1994.sp020244. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Perrins R, Roberts A. Cholinergic and electrical synapses between synergistic spinal motoneurones in the Xenopus laevis embryo. J Physiol. 1995;485:135–144. doi: 10.1113/jphysiol.1995.sp020718. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pfaffinger PJ, Martin JM, Hunter DD, Nathanson NM, Hille B. GTP-binding proteins couple cardiac muscarinic receptors to a K channel. Nature. 1985;317:536–538. doi: 10.1038/317536a0. [DOI] [PubMed] [Google Scholar]
- Quinlan KA, Placas PG, Buchanan JT. Cholinergic modulation of the locomotor network in the lamprey spinal cord. J Neurophysiol. 2004;92:1536–1548. doi: 10.1152/jn.01053.2003. [DOI] [PubMed] [Google Scholar]
- Rekling JC, Funk GD, Bayliss DA, Dong XW, Feldman JL. Synaptic control of motoneuronal excitability. Physiol Rev. 2000;80:767–852. doi: 10.1152/physrev.2000.80.2.767. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Russier M, Kopysova IL, Ankri N, Ferrand N, Debanne D. GABA and glycine co-release optimizes functional inhibition in rat brainstem motoneurons in vitro. J Physiol. 2002;541:123–137. doi: 10.1113/jphysiol.2001.016063. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Russo RE, Hounsgaard J. Dynamics of intrinsic electrophysiological properties in spinal cord neurones. Prog Biophys Mol Biol. 1999;72:329–365. doi: 10.1016/s0079-6107(99)00011-5. [DOI] [PubMed] [Google Scholar]
- Sah P. Ca2+-activated K+ currents in neurones: types, physiological roles and modulation. Trends Neurosci. 1996;19:150–154. doi: 10.1016/s0166-2236(96)80026-9. [DOI] [PubMed] [Google Scholar]
- Sakmann B, Noma A, Trautwein W. Acetylcholine activation of single muscarinic K+ channels in isolated pacemaker cells of the mammalian heart. Nature. 1983;303:250–253. doi: 10.1038/303250a0. [DOI] [PubMed] [Google Scholar]
- Schafer MK, Eiden LE, Weihe E. Cholinergic neurons and terminal fields revealed by immunohistochemistry for the vesicular acetylcholine transporter. II. The peripheral nervous system. Neuroscience. 1998;84:361–376. doi: 10.1016/s0306-4522(97)80196-0. [DOI] [PubMed] [Google Scholar]
- Schnee ME, Brown BS. Selectivity of linopirdine (DuP 996), a neurotransmitter release enhancer, in blocking voltage-dependent and calcium-activated potassium currents in hippocampal neurons. J Pharmacol Exp Ther. 1998;286:709–717. [PubMed] [Google Scholar]
- Seeger T, Alzheimer C. Muscarinic activation of inwardly rectifying K+ conductance reduces EPSPs in rat hippocampal CA1 pyramidal cells. J Physiol. 2001;535:383–396. doi: 10.1111/j.1469-7793.2001.00383.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seutin V, Scuvee-Moreau J, Dresse A. Evidence for a non-GABAergic action of quaternary salts of bicuculline on dopaminergic neurones. Neuropharmacology. 1997;36:1653–1657. doi: 10.1016/s0028-3908(97)00147-0. [DOI] [PubMed] [Google Scholar]
- Smith JC, Feldman JL, Schmidt BJ. Neural mechanisms generating locomotion studied in mammalian brain stem-spinal cord in vitro. FASEB J. 1988;2:2283–2288. doi: 10.1096/fasebj.2.7.2450802. [DOI] [PubMed] [Google Scholar]
- Sodickson DL, Bean BP. GABAB receptor-activated inwardly rectifying potassium current in dissociated hippocampal CA3 neurons. J Neurosci. 1996;16:6374–6385. doi: 10.1523/JNEUROSCI.16-20-06374.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Storm JF. An after-hyperpolarization of medium duration in rat hippocampal pyramidal cells. J Physiol. 1989;409:171–190. doi: 10.1113/jphysiol.1989.sp017491. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Surprenant A, North RA. Mechanism of synaptic inhibition by noradrenaline acting at alpha 2-adrenoceptors. Proc R Soc Lond B Biol Sci. 1988;234:85–114. doi: 10.1098/rspb.1988.0039. [DOI] [PubMed] [Google Scholar]
- Svirskis G, Hounsgaard J. Depolarization-induced facilitation of a plateau-generating current in ventral horn neurons in the turtle spinal cord. J Neurophysiol. 1997;78:1740–1742. doi: 10.1152/jn.1997.78.3.1740. [DOI] [PubMed] [Google Scholar]
- Takahashi T. Inward rectification in neonatal rat spinal motoneurones. J Physiol. 1990;423:47–62. doi: 10.1113/jphysiol.1990.sp018010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takigawa T, Alzheimer C. G protein-activated inwardly rectifying K+ (GIRK) currents in dendrites of rat neocortical pyramidal cells. J Physiol. 1999;517:385–390. doi: 10.1111/j.1469-7793.1999.0385t.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takigawa T, Alzheimer C. Phasic and tonic attenuation of EPSPs by inward rectifier K+ channels in rat hippocampal pyramidal cells. J Physiol. 2002;539:67–75. doi: 10.1113/jphysiol.2001.012883. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Topert C, Doring F, Wischmeyer E, Karschin C, Brockhaus J, Ballanyi K, Derst C, Karschin A. Kir2.4: a novel K+ inward rectifier channel associated with motoneurons of cranial nerve nuclei. J Neurosci. 1998;18:4096–4105. doi: 10.1523/JNEUROSCI.18-11-04096.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vogalis F, Storm JF, Lancaster B. SK channels and the varieties of slow after-hyperpolarizations in neurons. Eur J Neurosci. 2003;18:3155–3166. doi: 10.1111/j.1460-9568.2003.03040.x. [DOI] [PubMed] [Google Scholar]
- Wang HS, McKinnon D. Potassium currents in rat prevertebral and paravertebral sympathetic neurones: control of firing properties. J Physiol. 1995;485:319–335. doi: 10.1113/jphysiol.1995.sp020732. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Welton J, Stewart W, Kerr R, Maxwell DJ. Differential expression of the muscarinic m2 acetylcholine receptor by small and large motoneurons of the rat spinal cord. Brain Res. 1999;817:215–219. doi: 10.1016/s0006-8993(98)01208-6. [DOI] [PubMed] [Google Scholar]
- Wheatley M, Stein RB. An in vitro preparation of the mudpuppy for simultaneous intracellular and electromyographic recording during locomotion. J Neurosci Meth. 1992;42:129–137. doi: 10.1016/0165-0270(92)90143-2. [DOI] [PubMed] [Google Scholar]
- Williams SR, Christensen SR, Stuart GJ, Hausser M. Membrane potential bistability is controlled by the hyperpolarization-activated current IH in rat cerebellar Purkinje neurons in vitro. J Physiol. 2002;539:469–483. doi: 10.1113/jphysiol.2001.013136. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Williams SR, Turner JP, Hughes SW, Crunelli V. On the nature of anomalous rectification in thalamocortical neurones of the cat ventrobasal thalamus in vitro. J Physiol. 1997;505:727–747. doi: 10.1111/j.1469-7793.1997.727ba.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilson DM, Larimer JL. The catch property of ordinary muscle. Proc Natl Acad Sci U S A. 1968;61:909–916. doi: 10.1073/pnas.61.3.909. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilson JM, Rempel J, Brownstone RM. Postnatal development of cholinergic synapses on mouse spinal motoneurons. J Comp Neurol. 2004;474:13–23. doi: 10.1002/cne.20089. [DOI] [PubMed] [Google Scholar]
- Xiang Z, Huguenard JR, Prince DA. Cholinergic switching within neocortical inhibitory networks. Science. 1998;281:985–988. doi: 10.1126/science.281.5379.985. [DOI] [PubMed] [Google Scholar]
- Zhang L, Krnjevic K. Apamin depresses selectively the after-hyperpolarization of cat spinal motoneurons. Neurosci Lett. 1987;74:58–62. doi: 10.1016/0304-3940(87)90051-6. [DOI] [PubMed] [Google Scholar]
- Zieglgansberger W, Reiter C. A cholinergic mechanism in the spinal cord of cats. Neuropharmacology. 1974;13:519–527. doi: 10.1016/0028-3908(74)90141-5. [DOI] [PubMed] [Google Scholar]