The hyperpolarisation-activated cation current (Ih) has been described in many vertebrate and invertebrate species and cell types. In neurons, Ih is involved in rhythmogenesis, membrane potential stabilisation and many other functions. In this work, we investigate the distribution and functional properties of Ih in identified leech neurons of intact segmental ganglia. We found Ih in the mechanosensory touch (T), pressure (P) and noxious (N) neurons, as well as in Retzius neurons. The current displayed its largest amplitude in P neurons and we investigated its biophysical and pharmacological properties in these cells. Ih was half-maximally activated at –65 mV and fully activated at –100 mV. The current mutually depended on both Na+ and K+ with a permeability ratio pNa/pK of ∼0.21. The reversal potential was approximately –35 mV. The time course of activation could be approximated by a single time constant of ∼370 ms at –60 mV, but required two time constants at –80 mV of ∼80 and ∼560 ms. The current was half-maximally blocked by 0.3 mmol l–1 Cs+ but was insensitive to the bradycardic agent ZD7288. The physiological function of this channel could be a subtle alteration of the firing behaviour of mechanosensory neurons as well as a stabilisation of the resting membrane potential.

The hyperpolarisation-activated cation current (Ih) is involved in a range of functions of cells across species (Biel et al., 2009; Pape, 1996). In neurons, the current allows the control of rhythmic activity [e.g. in the thalamocortical system (McCormick and Pape, 1990b)], plays a key role in determining and stabilising the resting membrane potential (Em) (Ludwig et al., 2003; Meuth et al., 2006; Pape, 1996) and has a wide range of other functions in processes such as dendritic integration, synaptic plasticity, synaptic transmission and the temporal processing of visual signals in the retina (for a review, see Biel et al., 2009). The current has several characteristic properties: (1) it is carried by both Na+ and K+ and has a reversal potential (Erev) of approximately –20 mV under physiological conditions, hence it is inwardly directed at rest, causing a membrane depolarisation upon activation; (2) it is activated by hyperpolarising voltage deflections negative to approximately –55 mV; (3) its activation is facilitated by cyclic adenosine monophosphate (cAMP); and (4) it is blocked by low concentrations of Cs+ and by the bradycardic agent ZD7288.

The first systematic description of a hyperpolarisation-activated cation conductance in leech neurons by Arbas and Calabrese showed that the electrical properties of heart interneurons (HN cells) include a Cs+-sensitive component that is vital for these cells' oscillatory behaviour (Arbas and Calabrese, 1987a; Arbas and Calabrese, 1987b). This component appeared as a slow partial repolarisation after hyperpolarising current injections (‘voltage sag’, see Fig. 1). Later, the underlying current was shown to have the typical properties of neuronal Ih found in mammals, in that it was fully activated near –70 to –80 mV, its activation kinetics were voltage dependent between –100 and –60 mV (with time constants of 700 ms and 2 s, respectively) and it was carried by both Na+ and K+ (Angstadt and Calabrese, 1989).

In leech HN cells, Ih helps to control heartbeat: a series of inhibitory postsynaptic potentials hyperpolarises the membrane and causes a slow activation of Ih, which, in turn, mediates the depolarising phase (Angstadt and Calabrese, 1989). This function is similar to that in the thalamic neurons of mammals. Here, the current can provide pacemaker depolarisations after the preceding activity of hyperpolarising currents, which then trigger depolarisation-activated conductances such as low-threshold Ca2+ channels (Akasu et al., 1993; McCormick and Pape, 1990a; McCormick and Pape, 1990b; Pape and McCormick, 1989). Another function is to stabilise neurons near their Em by lowering the input resistance upon hyperpolarising stimuli, causing both reduced amplitude of postsynaptic potentials (‘shunting inhibition’) and a return to more positive potentials. Ih was reported to play a role in stabilising the resting membrane potential in thalamic relay neurons, nodose sensory neurons and hippocampal interneurons (Doan and Kunze, 1999; Lupica et al., 2001; Meuth et al., 2006). Such a role has yet to be described for leech neurons, and a detailed analysis of Ih across leech neurons is not yet available.

Besides HN cells, a significant Ih that activates near –70 mV was also found in cultured leech Retzius neurons (Angstadt, 1999). In Retzius neurons of intact ganglia, voltage sags appeared to be substantially smaller, but Ih was nevertheless thought to be involved in the control of spontaneous action potential activity (Angstadt, 1999; Coulon et al., 2008). Marked voltage sags have been described in pressure (P) neurons in situ (Jansen and Nicholls, 1973; Klees et al., 2005), but the underlying mechanism has not been investigated. Ih was also found in salivary glands of the giant Amazon leech, Haementeria ghilianii (Wuttke and Berry, 1992).

The aims of this study were to: (1) determine which neurons show Ih or the characteristic voltage sag; (2) characterise key properties such as activation kinetics, voltage dependence, effects of impermeable cations, pharmacological profile and ion fluxes that constitute the current; and (3) elucidate the possible functional role this current plays in leech P neurons by assessing the excitability and the effects on the time course of AP firing.

We found that Ih preferentially appears in leech mechanosensory neurons and is largest in P neurons. We concentrated on P neurons for the biophysical and pharmacological characterisation of the current properties and for a discussion of the current's functional role.

Preparation

Preparation was done as described previously (Coulon et al., 2008). Leeches were purchased from a commercial supplier (Zaug GmbH, Biebertal, Germany), or taken from the laboratory's own breeding stock, and identified as Hirudo verbana Carena 1820 (Annelida, Hirudinea) (see Siddall et al., 2007). Intact segmental ganglia were continuously superfused at a rate of ∼5 ml min–1, exchanging the chamber volume (0.05 ml) approximately 100 times per minute.

Electrophysiological recordings

Neurons were impaled by two sharp electrolyte-filled microelectrodes for recording Em and for current injection. The electrodes were pulled on a vertical puller (Narishige PE-2, London, UK) from borosilicate capillaries (outer/inner diameter: 1.5 mm/0.86 mm, 0.15 mm filament; GC 150F-15, Clark Electromedical Instruments, Pangbourne, UK) and filled with 0.5 mol l–1 K2SO4 and 5 mmol l–1 KCl. Electrode resistance was ∼90 MΩ. Electrodes were connected to a custom-made two-electrode voltage-clamp amplifier (modified TEC-05L, NPI Instruments, Tamm, Germany). The bath was grounded via an agar bridge. The output signals were digitised by an A/D converter (custom built, Zentralwerkstatt Biologie, Heinrich-Heine-Universität Düsseldorf, Germany; or Digidata 1322A, Molecular Devices, Sunnyvale, CA, USA) and were saved and analysed on a computer running in-house acquisition software (Eberhard von Berg, Institut für Neurobiologie, Heinrich-Heine-Universität Düsseldorf) or pClamp (Molecular Devices). Sampling frequency was 1 kHz for slow signals or 10 kHz for fast signals. Stimulus parameters were programmed on a MAX 21 pulse generator (Zeitz Instruments, Augsburg, Germany). The experimental setup was shielded by a Faraday cage and dampened by shock absorbers. Recordings were performed at room temperature (21°C). Fast current oscillations lasting roughly 10 ms after holding potential deflections in the voltage-clamp mode are not shown in the current traces to improve clarity.

Identification of neurons

The 21 segmental ganglia contain ∼400 neurons arranged in six packets. The neurons investigated in this work can be divided into neurosecretory (Retzius and Leydig), motor [annulus erector (AE) and anterior pagoda (AP)] and mechanosensory [touch (T), pressure (P) and noxious (N)]. When ganglia are fixed ventral side up, Retzius neurons are easily identifiable based on their large size (∼80 μm diameter of the soma) and central position. In the adjacent, posterior-lateral packet there are two pairs of P neurons (P1, P2), which can be distinguished by their electrophysiological characteristics. Depolarising current injections evoke large-amplitude action potentials and hyperpolarising current injections evoke characteristic voltage sags (see Fig. 1). The two anterior-lateral packets contain N, T and AP neurons. N and AP neurons are similar in size to P neurons, but N neurons generate very large action potentials with a prominent afterhyperpolarisation immediately after microelectrode impalement. AP neurons show spontaneous action potentials with low amplitude (∼10 mV) at a frequency of ∼5 Hz and a characteristic ‘pagoda’ shape (Nicholls, 1987). T neurons are smaller in size and do not show spontaneous action potentials. AE neurons are situated adjacent to neuron 251 and generate spontaneous low-amplitude (∼7 mV) action potentials at ∼10 Hz. Leydig neurons are adjacent to P2 neurons, smaller in size and fire spontaneously at less than 4 Hz (Arbas and Calabrese, 1990).

Bath solutions and drugs

The standard leech saline (SLS) used for bath perfusion had the following composition (in mmol l–1): 85 NaCl, 4 KCl, 2 CaCl2, 1 MgCl2 and 10 HEPES. The pH was adjusted to 7.40 with 1 mol l–1 NaOH, which increased the Na+ concentration by 4 mmol l–1. The osmolality of the SLS was 190 mosmol kg–1 H2O (Osmomat 030, Gonotec, Berlin, Germany). Na+-free solution was prepared by substitution of Na+ with N-methyl-d-glucammonium (NMDG+, prepared from N-methyl-d-glucamine and HCl; Sigma-Aldrich, Munich, Germany). Because of differences in osmotic activity, 105 mmol l–1 NMDG-Cl was used to substitute 89 mmol l–1 NaCl. In solutions containing 89 mmol l–1 Li+, Na+ was omitted and the pH was adjusted using LiOH. Finally, in solutions with varying K+ concentrations, K+ was either removed, or added to the solution without osmotic compensation. Monovalent and polyvalent cations were added as chloride salts without osmotic compensation.

In most preparations, the activation curve of Ih can be shifted towards more positive potentials by elevations in the intracellular cAMP concentration (Pape, 1996), thus increasing active Ih under resting conditions. In order to determine whether leech Ih channels are also sensitive to cAMP, we used the membrane-permeable cAMP analogue dibutyryl-cAMP (dbcAMP). Similarly, we applied 3-isobutyl-1-methylxanthine (IBMX, Sigma-Aldrich) to inhibit phosphodiesterases and, thus, increase the intracellular cAMP concentration (Bobker and Williams, 1989). To directly block Ih, we used Cs+ (CsCl, Sigma-Aldrich) or ZD7288 (4-ethylphenylamino-1,2-dimethyl-6-methylaminopyrimidinium chloride; Tocris, Bristol, UK). For experiments in the presence of ZD7288, the bath perfusion was stopped after adding ZD7288 to the SLS.

Calculation of ion selectivity and analysis of experimental data

Ih is carried by cation channels that are selective for Na+ and K+ (Ho et al., 1994). Determining Erev for Ih allows the calculation of the permeability ratio (pNa/pK) by rearranging Goldman's equation (see Aidley, 1989):
formula
(1)
where [K+]o and [Na+]o are the extracellular and [K+]i and [Na]+i the intracellular concentrations of K+ and Na+, respectively, z is the number of elementary charges transferred, R is the universal gas constant, T is the absolute temperature and Erev is the reversal potential. The conductance ratio (gNa/gK) can be calculated using the formalism from Hodgkin and Huxley (Hodgkin and Huxley, 1952) (see Aidley, 1989):
formula
(2)
where EK and ENa are the equilibrium potentials for K+ and Na+, respectively. At a given Em, Ih can be calculated by (Aidley, 1989):
formula
(3)
By combining Eqns 2 and 3, gNa and gK, as well as the currents carried by Na+ and K+, can be calculated. For the intracellular and extracellular concentrations of Na+ and K+, data obtained with ion-sensitive microelectrodes were used (see below); these data were also used to calculate the equilibrium potentials of Na+ and K+.

Experimental data was analysed and fitted using Origin Software (Microcal, Northampton, MA, USA). Boltzmann's equation was used to calculate the voltage sensitivity of Ih activation, Hill's equation was used to calculate the dose–response curve, and single or sums of e-functions were used to describe the activation and deactivation kinetics of the current.

Voltage sag in identified leech neurons

The injection of a negative current into Retzius neurons or into the mechanosensory T, P and N neurons induced a membrane hyperpolarisation that was maximal after approximately 50 ms, but subsequently became smaller. The Em reached a less negative plateau a few hundred milliseconds later (Fig. 1). This voltage sag was virtually abolished after the addition of 1–5 mmol l–1 Cs+ to the bath (see Fig. 11A), indicating that it was mediated by Ih channels. The amplitude of the voltage sag was determined as the depolarisation between the peak of the negative voltage deflection and the potential at the end of a given current injection (Table 1). A voltage sag was absent in Leydig, AE and AP neurons (see Fig. 1). The voltage sag was largest by far in P neurons (see Fig. 1, Table 1), prominent in N neurons and small in Retzius neurons, where a voltage sag was discernible in the majority of cells (five out of nine). The smallest voltage sag was observed in three out of four tested T neurons.

Table 1.

Size of the voltage sag in touch (T), pressure (P), noxious (N) and Retzius neurons

Size of the voltage sag in touch (T), pressure (P), noxious (N) and Retzius neurons
Size of the voltage sag in touch (T), pressure (P), noxious (N) and Retzius neurons

The functional properties of the Ih channels were investigated in more detail in P neurons, because, in these cells, the voltage sag was largest. Furthermore, P neurons are sufficiently large and robust to routinely and reliably allow the application of the two-electrode voltage-clamp technique.

Ih in leech P neurons

A typical voltage-clamp experiment showing the activation of Ih in a P neuron is presented in Fig. 2. Initially, the Em of the cell was clamped to a reference potential of –50 mV, which was slightly more negative than the mean Em of leech P neurons (–44.3±4.9 mV, N=67) (see Schlue and Deitmer, 1984). This required a holding current of –0.4±0.3 nA (N=40). Clamping the neuron to –80 mV induced an inward current (‘1’ in Fig. 2B), reflecting the net change in membrane currents through ion channels that were already active at –50 mV (instantaneous current, Ii). Subsequently, an additional current mediated by Ih channels developed, which was evident from its biophysical and pharmacological properties (see below). Ih was maximal after approximately 0.5 s and persisted throughout the hyperpolarising pulse, indicating that the channels did not inactivate. Provided that the Ii remained constant during the hyperpolarising pulse, the Ih that was activated when stepping from –50 to –80 mV was obtained by subtracting Ii from the current at the end of the pulse (steady-state or slow inward current, ISS; ‘2’ in Fig. 2B). After returning to –50 mV, an inwardly directed tail current (‘3’ in Fig. 2B) remained, which slowly declined as the Ih channels deactivated (‘4’ in Fig. 2B).

Fig. 1.

Voltage sags in identified leech neurons. Upon injection of a hyperpolarising current of the given amplitude, Retzius neurons as well as the mechanosensory touch (T), pressure (P) and noxious (N) neurons showed voltage sags; i.e. after reaching a maximum hyperpolarisation, the membrane potential (Em) repolarised partially. After cessation of the injection current, a transient depolarisation occurred, which in P neurons sometimes caused the generation of one or more action potentials. In Leydig, annulus erector (AE) and anterior pagoda (AP) neurons, voltage sags were not observed. The injection current (–2 to –6 nA for 500 or 800 ms) was adjusted such that the cells initially hyperpolarised to approximately –100 mV. Note the occurrence of spontaneous action potentials in Leydig, AE and AP neurons. The action potentials of P and Leydig neurons are truncated.

Fig. 1.

Voltage sags in identified leech neurons. Upon injection of a hyperpolarising current of the given amplitude, Retzius neurons as well as the mechanosensory touch (T), pressure (P) and noxious (N) neurons showed voltage sags; i.e. after reaching a maximum hyperpolarisation, the membrane potential (Em) repolarised partially. After cessation of the injection current, a transient depolarisation occurred, which in P neurons sometimes caused the generation of one or more action potentials. In Leydig, annulus erector (AE) and anterior pagoda (AP) neurons, voltage sags were not observed. The injection current (–2 to –6 nA for 500 or 800 ms) was adjusted such that the cells initially hyperpolarised to approximately –100 mV. Note the occurrence of spontaneous action potentials in Leydig, AE and AP neurons. The action potentials of P and Leydig neurons are truncated.

Fig. 2.

Hyperpolarisation-activated cation (Ih) channels in a leech P neuron. Shifting Em from the reference potential of –50 to –80 mV (A) induced the slow activation of a persistent inward current (B), which is mediated by Ih channels, as evident from the biophysical and pharmacological properties. The amplitude of Ih is given by the difference between the instantaneous current at the beginning of the voltage step (Ii; 1) and the membrane current at its end (steady-state or slow inward current, ISS; 2). After return to the reference potential, a slowly falling tail current remained (3), which reflects the deactivation of the Ih channels. After complete deactivation, the current returned to resting values (4). The holding current at –50 mV was –0.4±0.3 nA (N=40).

Fig. 2.

Hyperpolarisation-activated cation (Ih) channels in a leech P neuron. Shifting Em from the reference potential of –50 to –80 mV (A) induced the slow activation of a persistent inward current (B), which is mediated by Ih channels, as evident from the biophysical and pharmacological properties. The amplitude of Ih is given by the difference between the instantaneous current at the beginning of the voltage step (Ii; 1) and the membrane current at its end (steady-state or slow inward current, ISS; 2). After return to the reference potential, a slowly falling tail current remained (3), which reflects the deactivation of the Ih channels. After complete deactivation, the current returned to resting values (4). The holding current at –50 mV was –0.4±0.3 nA (N=40).

The activation of Ih was evident at Em values slightly more negative than –50 mV (Fig. 3). With increasing hyperpolarisation, Ih became more and more prominent. At –80 mV, it was approximately as large as Ii (Fig. 3B, Fig. 4B). The tail current saturated with increasing, preceding hyperpolarisation, indicating that Ih channel activation was all but complete at approximately –80 mV (see Fig. 5).

The dependence of the Ii on the Em was linear, the slope corresponding to an average input resistance of 23 MΩ (Fig. 3B, red line). In contrast, the membrane current at the end of the hyperpolarising pulse increased over-proportionally because of the increasing activation of Ih channels. At Em values more negative than –70 mV, the slope of the current–voltage relationship appeared to approach a constant value, corresponding to an input resistance of 9 MΩ. At still more negative potentials the slope became smaller again, corresponding to an input resistance of 14 MΩ (see Fig. 8B, black squares).

Kinetics of activation and deactivation

The speed of Ih activation increased with growing membrane hyperpolarisation, as was previously described for other preparations (e.g. Budde et al., 2008) (see Fig. 3A). We approximated the activation kinetics by exponential functions as illustrated in Fig. 4, which shows the membrane currents induced by hyperpolarising the cell membrane to –60 or –80 mV. When the time course of Ih activation was fitted by a single exponential function, time constants of 368±74 ms (for –60 mV) and 158±42 ms (for –80 mV) were obtained (N=12). At an Em of –60 mV, the recorded trace and calculated curve corresponded very well to each other, but deviations were evident at –80 mV (see Fig. 4B, red trace). At this holding potential, a good fit was obtained by the sum of two exponential functions with time constants of 78±22 and 557±84 ms, whereby the fast component contributed approximately 75% to the total Ih. The decay of the tail current could be approximated by a single exponential function, with time constants ranging from 150 to 230 ms, and it was independent of the strength of the preceding hyperpolarisation (data not shown). In some recordings, the activation of Ih appeared to follow a slightly sigmoidal time course (see traces in Fig. 2B, Fig. 3A, Fig. 4B; see Discussion).

Fig. 3.

Voltage dependence of Ii and ISS. (A) Membrane currents upon shifting Em from –50 to –80 mV (in 2 mV steps) for 2 s, recorded at intervals of 30 s. With increasing hyperpolarisation, the activation of Ih became progressively faster, while the decay of the tail current appeared to follow an invariant time course. (B) Relationship between membrane current and Em, either at the beginning of the hyperpolarising step (Ii, open circles) or at its end, after activation of the Ih channels (ISS, closed circles). Data points are means ± s.d. of 23 experiments. The voltage dependence of Ii was well fitted by a straight line (red line), the slope of which corresponded to an input resistance of 23 MΩ. In contrast, ISS increased over-proportionally because of the increasing activation of the Ih channels. The difference between the two current–voltage relationships (ISSIi) gives the amplitude of Ih activated by the respective voltage step.

Fig. 3.

Voltage dependence of Ii and ISS. (A) Membrane currents upon shifting Em from –50 to –80 mV (in 2 mV steps) for 2 s, recorded at intervals of 30 s. With increasing hyperpolarisation, the activation of Ih became progressively faster, while the decay of the tail current appeared to follow an invariant time course. (B) Relationship between membrane current and Em, either at the beginning of the hyperpolarising step (Ii, open circles) or at its end, after activation of the Ih channels (ISS, closed circles). Data points are means ± s.d. of 23 experiments. The voltage dependence of Ii was well fitted by a straight line (red line), the slope of which corresponded to an input resistance of 23 MΩ. In contrast, ISS increased over-proportionally because of the increasing activation of the Ih channels. The difference between the two current–voltage relationships (ISSIi) gives the amplitude of Ih activated by the respective voltage step.

Fig. 4.

Kinetics of Ih activation and deactivation. (A) At moderate hyperpolarisations (e.g. –60 mV), the activation of Ih was well fitted by a single exponential function. The upper trace shows the membrane current upon a hyperpolarisation from –50 to –60 mV (black trace), together with the approximation of Ih by a single exponential function with τ=368 ms and an amplitude of –0.38 nA (red curve). The lower trace shows the difference between recorded trace and calculated curve at an increased magnification, allowing an assessment of fit quality. See B for scale bars. (B) At stronger hyperpolarisations, the approximation of Ih by a single exponential function became progressively worse, but we obtained a satisfying fit using a sum of two exponential functions. The black trace shows the membrane current upon a hyperpolarisation to –80 mV, together with the approximation by a single exponential function (τ=158 ms; red curve) or the sum of two exponential functions (τ1=78 ms, τ2=557 ms; green curve). The fast component contributed –0.93 nA to the total current and the slow component contributed –0.33 nA. The lower two traces show the difference between the recorded trace and the curves calculated with one exponential function (red) or the sum of two exponential functions (green). The decline of the tail currents in A and B could be reasonably fitted by a single exponential function with a time constant of approximately 200 ms.

Fig. 4.

Kinetics of Ih activation and deactivation. (A) At moderate hyperpolarisations (e.g. –60 mV), the activation of Ih was well fitted by a single exponential function. The upper trace shows the membrane current upon a hyperpolarisation from –50 to –60 mV (black trace), together with the approximation of Ih by a single exponential function with τ=368 ms and an amplitude of –0.38 nA (red curve). The lower trace shows the difference between recorded trace and calculated curve at an increased magnification, allowing an assessment of fit quality. See B for scale bars. (B) At stronger hyperpolarisations, the approximation of Ih by a single exponential function became progressively worse, but we obtained a satisfying fit using a sum of two exponential functions. The black trace shows the membrane current upon a hyperpolarisation to –80 mV, together with the approximation by a single exponential function (τ=158 ms; red curve) or the sum of two exponential functions (τ1=78 ms, τ2=557 ms; green curve). The fast component contributed –0.93 nA to the total current and the slow component contributed –0.33 nA. The lower two traces show the difference between the recorded trace and the curves calculated with one exponential function (red) or the sum of two exponential functions (green). The decline of the tail currents in A and B could be reasonably fitted by a single exponential function with a time constant of approximately 200 ms.

Voltage dependence of Ih channel activity

The tail current was recorded at constant electromotive forces, so that its amplitude should be proportional to the activity of the Ih channels at the end of the hyperpolarising pulse. The dependence of the tail current on the preceding Em is shown in Fig. 5. An approximation of the experimental data by a Boltzmann function gave a half-maximal activation of the Ih channels at –65 mV. According to this approximation, the activation of Ih channels started at –30 mV, and at –100 mV the channels were fully activated. The data in Fig. 5 also illustrate the large variability of the tail current in individual cells. Following hyperpolarisation to –80 mV, the tail current amplitudes ranged between –0.1 and –0.8 nA.

Fig. 5.

Dependence of Ih channel activity on Em. Relationship between the tail current recorded at –50 mV and Em during the preceding hyperpolarisation. Data were obtained from 21 P neurons. In seven cells, Em was shifted in 5 mV steps up to –120 mV and, in the remaining 14 cells, in 2 mV steps up to –80 mV (pulse duration 2 s, pulse interval 30 s). The continuous line represents an approximation of the data by a Boltzmann function, according to which the Ih channels are half-maximally activated at Em=–65 mV.

Fig. 5.

Dependence of Ih channel activity on Em. Relationship between the tail current recorded at –50 mV and Em during the preceding hyperpolarisation. Data were obtained from 21 P neurons. In seven cells, Em was shifted in 5 mV steps up to –120 mV and, in the remaining 14 cells, in 2 mV steps up to –80 mV (pulse duration 2 s, pulse interval 30 s). The continuous line represents an approximation of the data by a Boltzmann function, according to which the Ih channels are half-maximally activated at Em=–65 mV.

Erev and ion permeability

The Erev of Ih was determined by analysing the current flow through fully activated Ih channels at different Em (Fig. 6A). For this purpose, Em was shifted very briefly to a defined test potential and Ii at this potential was determined. Then, after a short pause, the cells were hyperpolarised to –100 mV for 3 s to achieve complete Ih channel activation. Finally, Em was clamped again to the test potential. The membrane current measured at this time point is the sum of the Ii and the current mediated by the fully activated Ih channels. Its amplitude ideally depends on electromotive driving forces only.

The dependence of the maximal Ih on Em is shown in Fig. 6B. Linear fits of the data gave a mean Erev of –35±17 mV (N=7). Using this value and data on the intracellular and extracellular concentrations of Na+ and K+ previously obtained in our laboratory with ion-sensitive microelectrodes, pNa/pK was calculated to be 0.21. The calculation is based on the following values: [K+]o=5.8 mmol l–1 (at a bath concentration of 4 mmol l–1 K+) (see Schlue and Deitmer, 1980), [K+]i=97 mmol l–1 (Schlue, 1991) (see also Klees et al., 2005; Schlue and Deitmer, 1984), [Na+]o=89 mmol l–1 (assumed to correspond to the Na+ concentration of the bath solution) and [Na+]i=7.5 mmol l–1 (Hintz, 1999; Lucht, 1998) (see also Schlue, 1991). The equilibrium potentials were calculated to ENa=+62 mV and EK=–72 mV, and from the electromotive forces at Erev we obtained a gNa/gK of 0.38. It could be argued that [K+]o should be identical to the bath. In this case, pNa/pK would be 0.25 and gNa/gK would be 0.47. However, when the ganglion capsule was intact, the extracellular potassium concentration in the nerve cell body region was determined to be 5.8±0.6 mmol l–1 (mean ± s.d., N=27) (Schlue and Deitmer, 1980), whereas it was identical to the bath only after opening the ganglion capsule. Because the ganglion capsule was always intact in our experiments, we assumed [K+]o was equal to 5.8 mmol l–1 for all subsequent calculations and discussions.

Fig. 6.

Reversal potential of Ih. (A) Experimental protocol. To determine the instantaneous membrane current (Ii) at a given test potential (here, –80 mV), Em was shifted correspondingly for 50 ms (upper trace). During this period, the membrane current increased slightly, probably because of the onset of Ih channel activation (lower trace). Ii was estimated by visual extrapolation of the recorded trace to the beginning of the voltage jump (dark grey circle). After a 2 s pause, Em was shifted to –100 mV for 3 s to achieve full activation of the Ih channels. Subsequently, Em was shifted back to the test potential, in order to determine the maximum Ih at this potential. For this, the previously recorded Ii was subtracted from the current recorded immediately after the potential jump from –100 mV to the test potential (light grey circle). This protocol was applied every 10 s at different test potentials to determine the membrane currents mediated by the fully activated Ih channels at –30 to –90 mV in 10 mV increments. (B) Current–voltage relationships. The data points represent the membrane currents mediated by the fully activated Ih channels at different Em, as recorded from seven P neurons. In two cells, a reversal of Ih was not observed; in one cell Ih could only be measured between –90 and –60 mV. An error-weighted linear fit of the averaged data gave an Erev of Ih of –35 mV (N=7, red line).

Fig. 6.

Reversal potential of Ih. (A) Experimental protocol. To determine the instantaneous membrane current (Ii) at a given test potential (here, –80 mV), Em was shifted correspondingly for 50 ms (upper trace). During this period, the membrane current increased slightly, probably because of the onset of Ih channel activation (lower trace). Ii was estimated by visual extrapolation of the recorded trace to the beginning of the voltage jump (dark grey circle). After a 2 s pause, Em was shifted to –100 mV for 3 s to achieve full activation of the Ih channels. Subsequently, Em was shifted back to the test potential, in order to determine the maximum Ih at this potential. For this, the previously recorded Ii was subtracted from the current recorded immediately after the potential jump from –100 mV to the test potential (light grey circle). This protocol was applied every 10 s at different test potentials to determine the membrane currents mediated by the fully activated Ih channels at –30 to –90 mV in 10 mV increments. (B) Current–voltage relationships. The data points represent the membrane currents mediated by the fully activated Ih channels at different Em, as recorded from seven P neurons. In two cells, a reversal of Ih was not observed; in one cell Ih could only be measured between –90 and –60 mV. An error-weighted linear fit of the averaged data gave an Erev of Ih of –35 mV (N=7, red line).

Block by external Cs+ and insensitivity to ZD7288

A characteristic feature of Ih channels is the blockade by submillimolar concentrations of extracellular Cs+ (DiFrancesco, 1982; Kaupp and Seifert, 2001; Pape, 1996). In P neurons, Ih was half-maximally blocked at a Cs+ concentration of 0.3 mmol l–1. At 5 mmol l–1 Cs+, the block was virtually complete (Fig. 7). The onset of the block was fast and reversible within a few minutes. The time course of Ih activation was unaffected. The extracellular application of Ba2+, Co2+ or Ni2+ (5 mmol l–1 each), or of the K+ channel blocker tetraethylammonium (TEA, 20 mmol l–1), for up to 10 min had no effect on Ih (data not shown).

The bradycardic agent ZD7288 selectively blocks Ih channels in a variety of vertebrate and invertebrate preparations. The half-maximal effect is mostly exerted at concentrations of 10 to 50 μmol l–1 (see Gasparini and DiFrancesco, 1997; Pirtle et al., 2010; Shin et al., 2001). In P neurons, however, even concentrations of 1–10 mmol l–1, applied for up to 10 min, had no effect on Ih.

Fig. 7.

Ih is blocked by Cs+. (A) The presence of 2 mmol l–1 Cs+ reversibly suppressed Ih. Traces show the membrane current upon shifting Em from –50 to –80 mV for 1 s. Wash-in and wash-out time of Cs+ was 3 min, as indicated. (B) Ih in solutions containing various Cs+ concentrations were normalised to the mean current recorded before and after Cs+ application. The approximation of the data by using the Hill formalism (red line) gave a half-maximal block of Ih at a Cs+ concentration of 0.3 mmol l–1. The Hill coefficient was 0.98. Data points are means ± s.d. of four to six experiments.

Fig. 7.

Ih is blocked by Cs+. (A) The presence of 2 mmol l–1 Cs+ reversibly suppressed Ih. Traces show the membrane current upon shifting Em from –50 to –80 mV for 1 s. Wash-in and wash-out time of Cs+ was 3 min, as indicated. (B) Ih in solutions containing various Cs+ concentrations were normalised to the mean current recorded before and after Cs+ application. The approximation of the data by using the Hill formalism (red line) gave a half-maximal block of Ih at a Cs+ concentration of 0.3 mmol l–1. The Hill coefficient was 0.98. Data points are means ± s.d. of four to six experiments.

Fig. 8.

Ih in Na+-free solution. (A) Membrane currents upon shifting Em from –50 to –110 mV (in 5 mV steps) for 2 s, recorded at intervals of 30 s. The recordings were started 3 min after replacing extracellular Na+ by NMDG+. In Na+-free solution, Ih was virtually abolished. The holding current for clamping the cells to the reference potential was slightly reduced (–0.2±0.16 nA, N=7, cf. Fig. 2). (B) Relationship between the membrane current and Em, either at the beginning of the hyperpolarising step (Ii; open circles) or at its end, after activation of the Ih channels (ISS, Na+-free; closed circles). Data points are means ± s.d. of five experiments. The voltage dependence of Ii was fitted by a straight line, the slope of which indicated an input resistance of 27 MΩ (see Fig. 3). For comparison, the ISS recorded in SLS is also given (black squares; means ± s.d., N=7).

Fig. 8.

Ih in Na+-free solution. (A) Membrane currents upon shifting Em from –50 to –110 mV (in 5 mV steps) for 2 s, recorded at intervals of 30 s. The recordings were started 3 min after replacing extracellular Na+ by NMDG+. In Na+-free solution, Ih was virtually abolished. The holding current for clamping the cells to the reference potential was slightly reduced (–0.2±0.16 nA, N=7, cf. Fig. 2). (B) Relationship between the membrane current and Em, either at the beginning of the hyperpolarising step (Ii; open circles) or at its end, after activation of the Ih channels (ISS, Na+-free; closed circles). Data points are means ± s.d. of five experiments. The voltage dependence of Ii was fitted by a straight line, the slope of which indicated an input resistance of 27 MΩ (see Fig. 3). For comparison, the ISS recorded in SLS is also given (black squares; means ± s.d., N=7).

Ih in Na+-free solution

After replacing extracellular Na+ by NMDG+, Ih was virtually abolished (Fig. 8). At –100 mV, the mean Ih amplitude was –0.2 nA, as compared with –2.2 nA in SLS (Fig. 8B). The tail current was inverted and strongly reduced, with maximum amplitudes of approximately ∼0.1 nA (cf. Fig. 5); the time constant of deactivation varied between 250 and 400 ms. The voltage dependence of activation appeared to be unchanged: a Boltzmann approximation gave a half maximal activation at approximately –70 mV (data not shown). The effects of replacing extracellular Na+ by NMDG+ were fully reversible within a few minutes.

Replacing extracellular Na+ by Li+ also suppressed Ih and inverted the tail current (data not shown). The current for clamping the cells to –50 mV was increased (+1.3±0.6 nA, N=6). These effects occurred within 1 min after exchanging Na+ for Li+. Recovery was poor and, even 20 min after return to SLS, only a weak Ih was detectable. The tail current remained inverted and the current for clamping the cells to –50 mV often stayed increased.

Effect of extracellular K+

After reducing the K+ concentration of the bath solution, Ih and the tail current were strongly diminished (Fig. 9). Both currents were markedly increased when the extracellular K+ concentration was raised. At low K+ concentrations, Erev of Ih was shifted towards more negative values, corresponding to the shift in the K+ equilibrium potential. Under nominally K+-free conditions, it was measured to be –63±23 mV (N=12). Surprisingly, the Erev at increased bath K+ concentrations were also found to be more negative than in SLS (8 mmol l–1 K+: –40±5 mV, N=8; 16 mmol l–1: –39±13 mV, N=5).

Fig. 9.

Effect of extracellular K+ on Ih. (A) Superimposed membrane currents upon shifting Em from –50 to –80 mV for 2 s, recorded in the same P neuron 5 min after adjusting the K+ concentration in the bath as indicated. For better comparison, the traces were slightly shifted vertically to bring the instantaneous currents recorded immediately after shifting Em to congruence. (B) Amplitudes of Ih at –80 mV at different K+ concentrations in the bath solution. Data are means ± s.d. of five to 40 experiments (N-values are shown to the right of the error bars).

Fig. 9.

Effect of extracellular K+ on Ih. (A) Superimposed membrane currents upon shifting Em from –50 to –80 mV for 2 s, recorded in the same P neuron 5 min after adjusting the K+ concentration in the bath as indicated. For better comparison, the traces were slightly shifted vertically to bring the instantaneous currents recorded immediately after shifting Em to congruence. (B) Amplitudes of Ih at –80 mV at different K+ concentrations in the bath solution. Data are means ± s.d. of five to 40 experiments (N-values are shown to the right of the error bars).

Fig. 10.

Effect of cyclic nucleotides on Ih. (A,C) Ii (open symbols) and ISS (closed symbols) in the presence (circles) and absence (squares) of dibutyryl-cAMP (dbcAMP, 1 mmol l–1; A) or the phosphodiesterase blocker IBMX (1 mmol l–1; C). For clarity, error bars are only shown exemplarily. Although Ih was decreased by ∼50% in the presence of dbcAMP, the instantaneous current was unchanged. In IBMX, the amplitude of Ih remained unchanged but an overall decrease in the holding current was observed. (B,D) Relationship between tail currents at –50 mV and the preceding membrane hyperpolarisation, measured in the presence of dbcAMP (B) or IBMX (D). The recordings were started 10 min after adding dbcAMP or IBMX to the bath solution; single data points were measured at 10 s intervals. The continuous lines indicate approximations of the data by the Boltzmann equation, according to which the Ih channels were half-maximally activated at –66 mV (dbcAMP) or –67 mV (IBMX). Note that in the presence of IBMX, the maximum tail current was similar to that in SLS, whereas it was reduced by more than 50% in the presence of dbcAMP (cf. Fig. 5).

Fig. 10.

Effect of cyclic nucleotides on Ih. (A,C) Ii (open symbols) and ISS (closed symbols) in the presence (circles) and absence (squares) of dibutyryl-cAMP (dbcAMP, 1 mmol l–1; A) or the phosphodiesterase blocker IBMX (1 mmol l–1; C). For clarity, error bars are only shown exemplarily. Although Ih was decreased by ∼50% in the presence of dbcAMP, the instantaneous current was unchanged. In IBMX, the amplitude of Ih remained unchanged but an overall decrease in the holding current was observed. (B,D) Relationship between tail currents at –50 mV and the preceding membrane hyperpolarisation, measured in the presence of dbcAMP (B) or IBMX (D). The recordings were started 10 min after adding dbcAMP or IBMX to the bath solution; single data points were measured at 10 s intervals. The continuous lines indicate approximations of the data by the Boltzmann equation, according to which the Ih channels were half-maximally activated at –66 mV (dbcAMP) or –67 mV (IBMX). Note that in the presence of IBMX, the maximum tail current was similar to that in SLS, whereas it was reduced by more than 50% in the presence of dbcAMP (cf. Fig. 5).

Effects of cAMP and IBMX

In most cells, cAMP shifts the voltage range of Ih channel activation to more positive values (see Kaupp and Seifert, 2001). In the presence of the membrane-permeable cAMP analogue, dbcAMP, or the phosphodiesterase inhibitor, IBMX (each 1 mmol l–1), the Ih channels of leech P neurons were half-maximally activated at virtually the same Em as in SLS (Fig. 10). Amplitudes of Ih and tail current were reduced by approximately 50% upon dbcAMP application, whereas they were unchanged in the presence of IBMX. The instantaneous current was almost unaffected by both dbcAMP and IBMX.

Effect of Ih on membrane excitability

The activation of Ih channels reduced the depolarising current necessary to elicit an action potential (Fig. 11A), suggesting an increased excitability. Furthermore, approximately 30% of the cells generated action potentials without the injection of a depolarising current upon repolarisation from hyperpolarised potentials (see Fig. 1). These effects cannot be attributed to accommodation, because a membrane hyperpolarisation in the presence of millimolar concentrations of Cs+ never facilitated action-potential generation (Fig. 11A).

After activation of the Ih channels, the afterhyperpolarisation following an action potential shortened and subsequently reached a transient maximum (Fig. 11B, arrow). We never observed this when there was no preceding activation of the Ih channels. Ih channel activation did not affect amplitude and time course of the action potential. In the presence of Cs+, amplitude, afterhyperpolarisation and latency (stimulus to peak) of the action potentials were reduced, whereas the time course remained unchanged (Fig. 11C).

Ih in other leech neurons

When using the voltage-clamp technique to assess Ih in AP, Leydig, N, T and Retzius neurons, Ih could be excluded in Leydig but not in AP neurons. The maximum Ih current varied from cell type to cell type and occurred at voltages between –80 and –120 mV (Table 2). As Ii may already contain Ih, we additionally determined the difference between the steady-state currents under control conditions (i.e. with activated Ih) and in the presence of 2 mmol l–1 Cs+ (i.e. with Ih blocked; Cs+-sensitive current, ΔICs; Table 2). We found that the Cs+-sensitive current was consistently larger than the Ih current as determined from the difference between ISS and Ii. This corroborates that Ih is to some degree active at –50 mV, and/or that there is a rapidly activating Ih component, which is included in Ii, as has been reported previously (Budde et al., 2008; Mistrík et al., 2006; Proenza et al., 2002). When ΔICs is set in relation to the total holding current at a given holding potential, thus giving an estimate of the contribution of Ih to the overall membrane conductance, it becomes apparent that the mechanosensory N and P neurons have the largest Ih, followed by T, AP and Retzius neurons, whereas in Leydig neurons Ih appears to be virtually absent. We note that the values in Table 2 can only be a rough estimate, as Cs+ may inhibit conductances other than Ih.

Table 2.

Hyperpolarisation-activated cation currents (Ih) were observed in T, P, N, AP and Retzius neurons

Hyperpolarisation-activated cation currents (Ih) were observed in T, P, N, AP and Retzius neurons
Hyperpolarisation-activated cation currents (Ih) were observed in T, P, N, AP and Retzius neurons
Fig. 11.

Effect of Ih channel activation on membrane excitability and action potential waveform. (A) Membrane excitability. In this P neuron, action potentials could be reliably elicited by injecting a depolarising current of +5 nA for 5 ms, provided that the Ih channels had been activated by a preceding hyperpolarisation (–3 nA for 0.3 s; black trace). However, if the Ih channels were blocked by 1 mmol l–1 Cs+, action potentials were never generated in response to this stimulus (red trace). Each trace represents the superimposition of five single recordings to demonstrate the reproducibility of the waveforms. The effect of Cs+ was fully reversible within a few minutes. Identical results were obtained from five cells in total. (B) Action potential waveform. After activation of the Ih channels by a hyperpolarising current (–3 nA for 0.3 s), the afterhyperpolarisation following an action potential was shortened, and subsequently reached a transient maximum (red trace, arrow). This was never observed without preceding Ih channel activation (black trace). Amplitude and time course of the action potentials were not affected by Ih channel activation. As in A, each trace represents the superimposition of five single recordings. Traces were horizontally shifted to bring action potentials into congruence. In this P neuron, the injection of +2 nA for 5 ms was sufficient to reliably elicit an action potential, even without Ih channel activation. (C) Effect of Cs+ on action potentials. Superimposed traces of five consecutively elicited action potentials under control conditions (black) and in the presence of 1 mmol l–1 Cs+ (red). Traces were horizontally shifted to bring action potentials into congruence. Cs+ reduced the peak amplitudes and the afterhyperpolarisation. Moreover, the latency from the beginning of the current injection (arrows mark the stimulus artefact, +2 nA for 10 ms) to the peak of the action potential was also reduced. The effect of Cs+ was fully reversible. Similar results were obtained for five cells in total.

Fig. 11.

Effect of Ih channel activation on membrane excitability and action potential waveform. (A) Membrane excitability. In this P neuron, action potentials could be reliably elicited by injecting a depolarising current of +5 nA for 5 ms, provided that the Ih channels had been activated by a preceding hyperpolarisation (–3 nA for 0.3 s; black trace). However, if the Ih channels were blocked by 1 mmol l–1 Cs+, action potentials were never generated in response to this stimulus (red trace). Each trace represents the superimposition of five single recordings to demonstrate the reproducibility of the waveforms. The effect of Cs+ was fully reversible within a few minutes. Identical results were obtained from five cells in total. (B) Action potential waveform. After activation of the Ih channels by a hyperpolarising current (–3 nA for 0.3 s), the afterhyperpolarisation following an action potential was shortened, and subsequently reached a transient maximum (red trace, arrow). This was never observed without preceding Ih channel activation (black trace). Amplitude and time course of the action potentials were not affected by Ih channel activation. As in A, each trace represents the superimposition of five single recordings. Traces were horizontally shifted to bring action potentials into congruence. In this P neuron, the injection of +2 nA for 5 ms was sufficient to reliably elicit an action potential, even without Ih channel activation. (C) Effect of Cs+ on action potentials. Superimposed traces of five consecutively elicited action potentials under control conditions (black) and in the presence of 1 mmol l–1 Cs+ (red). Traces were horizontally shifted to bring action potentials into congruence. Cs+ reduced the peak amplitudes and the afterhyperpolarisation. Moreover, the latency from the beginning of the current injection (arrows mark the stimulus artefact, +2 nA for 10 ms) to the peak of the action potential was also reduced. The effect of Cs+ was fully reversible. Similar results were obtained for five cells in total.

For leech HN cells, and briefly for Retzius and P neurons, Ih has been described before (Angstadt, 1999; Angstadt and Calabrese, 1989; Arbas and Calabrese, 1987b; Coulon et al., 2008; Klees et al., 2005). However, a detailed description of the distribution and biophysical properties of leech Ih across cell types was lacking. The findings presented here can be summarised as follows. First, voltage sags and the underlying Ih were observed in T, P, N and Retzius neurons. Second, the current was largest in P neurons. Third, the biophysical properties of the current largely matched those described in other preparations: (a) half-maximum activation occurred at –65 mV, complete activation at –100 mV; (b) the activation could be fitted by a single time constant of ∼370 ms at –60 mV but required two time constants of ∼80 and ∼560 ms at –80 mV; (c) Erev was determined to be –35±17 mV; and (d) the current mutually depended on both extracellular Na+ and K+, with a pNa/pK of 0.21. Fourth, half-maximal block of Ih was achieved at a Cs+ concentration of 0.3 mmol l–1. The current was insensitive to ZD7288. Fifth, after elevation of the intracellular cAMP concentration, Ih in leech P neurons showed a blockade, rather than a voltage shift of the activation curve. Sixth, the channels caused bursts of APs when a neuron was relieved from hyperpolarisation and modulated the form of afterhyperpolarisations. And finally, the function of Ih seems to include a subtle alteration of firing behaviour in leech mechanosensory neurons as well as a stabilising effect on Em.

Activation kinetics

In some experiments, it was evident that Ih activation followed a sigmoidal time course (see Fig. 3A), as was found previously in leech heart interneurons [see fig. 2 in Angstadt and Calabrese (Angstadt and Calabrese, 1989)] and in most other systems (see Hestrin, 1987; Kaupp and Seifert, 2001). This phenomenon was interpreted as a delay-to-onset of the time-dependent component of Ih (Angstadt and Calabrese, 1989), i.e. the delay between the instantaneous component and the voltage-activated component of the membrane current. Unfortunately, in our experiments the early lag phase was short and mostly hidden by the initial oscillations in the holding current. When the early lag was ignored, the activation of Ih could be well approximated by a single exponential function, if Em was not too negative (Fig. 4A). At potentials that were more negative, a reasonable fit was only possible by applying a sum of two exponential functions (Fig. 4B). This biphasic activation has been reported previously (Budde et al., 1994; Pape, 1996) and was attributed to the existence of two distinct populations of Ih channels.

Ion selectivity

From the determined Erev of Ih (–35±17 mV, N=7; Fig. 6), pNa/pK was calculated to be 0.21 by using the Goldman equation. This calculation was based on intracellular and extracellular concentrations of Na+ and K+ measured previously by ion-sensitive microelectrodes. gNa/gK was determined to be 0.38. Comparison with permeability and conductance ratios obtained from other preparations reveals that the values calculated here are within the same range [pNa/pK between 0.2 and 0.4 (for a review, see Pape, 1996); gNa/gK between 0.25 and 0.67 (e.g. Bolívar et al., 2008; Macri et al., 2002)].

At –80 mV, Ih was on average approximately –1.3 nA (see Figs 2, 3). With this value and the above conductance ratio, gK is 21 nS and gNa is 8 nS. These values, in turn, allow us to calculate the tail current when returning from –80 mV to –50 mV to –0.44 nA, which corresponds nicely with the experimental data (see Fig. 5).

The suppression of Ih after replacing Na+ by Li+ shows that the channels are virtually impermeable to Li+, which is in accordance with previous findings (Biel et al., 2009; Pape, 1996; Wollmuth and Hille, 1992).

Pharmacological profile of Ih in leech neurons

One of the characteristics of Ih is its sensitivity to relatively low concentrations of Cs+ [0.1–5 mmol l–1 (Pape, 1996)]. Cs+ also blocks other voltage-dependent ion channels, such as delayed rectifier and inward rectifier K+ channels, complicating its use as a sole criterion for the identification of Ih. In leech P neurons, Ih was insensitive to Ni2+, Co2+ or Ba2+ (5 mmol l–1 each), as well as to TEA+ (20 mmol l–1; data not shown).

Another characteristic of Ih is its sensitivity to ZD7288. Leech P neurons were, however, completely insensitive to ZD7288 at concentrations as high as 10 mmol l–1. Nevertheless, in our view, most other experimental data shown here strongly argue in favour of the assumption that leech Ih channels are very similar, albeit not identical, to classical, ZD7288-sensitive Ih channels found in other preparations. Although it is beyond the scope of this work to elucidate why ZD7288 does not block the channels, it is not uncommon that known and established selective blockers are ineffective in the leech preparation in known concentration ranges (e.g. Johansen and Kleinhaus, 1986). Moreover, an ineffective blockade of leech Ih by ZD7288 was reported previously for HN cells (Tobin and Calabrese, 2005).

Leech Ih channels also show unusual behaviour concerning their reactions to cAMP. Ih channels are often regulated by cAMP such that they shift their voltage dependency of activation towards more positive potentials, whereas kinetics and amplitudes remain largely unaltered (Kaupp and Seifert, 2001; Pape, 1996; Pape and McCormick, 1989). In leech P neurons, the membrane-permeable analogue dbcAMP did not change the voltage dependency of activation, but instead caused a 50% reduction of the current. Lack of regulation by cAMP has been reported before and has been attributed to both HCN1 and HCN3 subunits of the channel, with HCN3 even being slightly inhibited by cAMP (Biel et al., 2009). In some systems that selectively express HCN1, cAMP shifted the activation curve by only a few millivolts (Santoro et al., 1998) or not at all (Shin et al., 2001). The insensitivity of the voltage dependence of activation to cAMP points to the expression of the HCN1 and HCN3 subtypes in leech neurons. In support of HCN1 expression is the comparably fast activation kinetics (Shin et al., 2001). In support of HCN3 expression is the inhibition by cAMP.

Multi-ion pores

The dependence of Ih on the presence of extracellular Na+ and K+ suggests that the fluxes of Na+ and K+ through the Ih channels are dependent on one another. Thus, after replacing extracellular Na+, Ih was all but absent. Under these conditions, a small Na+ efflux can be expected. However, this was not taken into account for the calculations, as the intracellular Na+ concentration drops in the absence of extracellular Na+ to values of 1 mmol l–1 and lower (Hintz, 1999), so that a Na+ efflux should be negligibly small. If, under these conditions, the K+ component of the current persisted, an inward current of –0.59 nA would be expected at –100 mV (gK=21 nS). However, the measured current was on average only –0.21 nA, indicating that gK was reduced to roughly one-third of its original value in the absence of extracellular Na+.

Ih was suppressed in the absence of K+ in the bath solution (Fig. 9), arguing in favour of a mutual dependency of the two ions on one another within the channel pore. In a K+-free bath solution, the extracellular K+ concentration in the vicinity of the neurons' somata was measured to be 1.6±0.6 mmol l–1 (Schlue and Deitmer, 1980). Under K+-free conditions, the lowest measured intracellular K+ was 78 mmol l–1 (Schlue, 1991; Schlue and Deitmer, 1984) and intracellular Na+ increased to ∼20 mmol l–1 (Lucht, 1998). With these ion concentrations and assuming that extracellular Na+ remains at 89 mmol l–1, Ih can be calculated with the given gK and gNa to –0.56 nA at –80 mV. However, the measured current was only –0.33±0.16 nA (N=12; Fig. 9B), suggesting that gNa was reduced by at least 40% in the K+-free bath solution. This mutual dependency of ion conductances within a channel pore is a characteristic feature of multi-ion pores (Hestrin, 1987; Hille, 2001).

Ih or the lack thereof in other neurons of the leech

Leydig neurons fire spontaneously at low rates [<4 Hz (Arbas and Calabrese, 1990)] and were found to regulate heart rate in the leech. We show here that Ih is all but absent in Leydig neurons, so that the intrinsic rhythm observed in these neurons cannot be mediated by Ih. The only non-mechanosensory neurons showing considerable Ih were Retzius neurons. In these cells, the current may affect spontaneous action-potential generation (Angstadt, 1999) and does not seem to be involved in volume regulation (Coulon, 2005) (see Coulon et al., 2008). The voltage sag and Ih were smallest in T neurons, and it is questionable whether such a small current can have a functional relevance.

Physiological relevance

The nervous system of the medicinal leech is organised in ganglia that contain various neurons with well-known functions. This offers the opportunity to correlate physiological and functional aspects. The hyperpolarisation-activated inward current of leech HN cells was slightly different in its biophysical properties when compared with the Ih described here in leech P neurons. It began to activate near –50 mV, became fully activated between –70 and –80 mV, and had slow and voltage-dependent activation kinetics with single time constants ranging from 2 s at –60 mV to 700 ms at –100 mV. It had an Erev of –21±5 mV and was blocked by 1–5 mmol l–1 extracellular Cs+ (Angstadt and Calabrese, 1989). HN cells interconnect by inhibitory chemical synapses, and their normal electrical activity consists of bursts of action potentials separated by periods of inhibition. The function of Ih in HN cells seems obvious: the cells hyperpolarise during the silent state so that Ih activates. This activation depolarises the membrane and helps the cells escape from the hyperpolarisation caused by inhibitory inputs. Accordingly, Cs+ slows (Masino and Calabrese, 2002; Tobin and Calabrese, 2005) or even abolishes oscillatory behaviour (Angstadt and Calabrese, 1989).

The neurons described in this work displaying Ih were mostly mechanosensory and not spontaneously active. Thus, a pacemaker function of Ih like that described for HN cells can be excluded. Instead, two further functions may exist in leech P neurons: (1) a stabilising effect on Em and (2) a subtle alteration of the firing behaviour. These neurons respond to mechanical stimuli of the skin with barrages of action potentials, proportional to the stimulus intensity, followed by a prominent hyperpolarisation that lasts for seconds or minutes and is caused by the activation of Ca2+-activated K+ channels and the activity of the Na+/K+-ATPase (Baylor and Nicholls, 1969; Jansen and Nicholls, 1973; Scuri et al., 2002). This hyperpolarisation will lead to the activation of Ih, which then helps to determine both the strength and duration of this hyperpolarisation. Moreover, a marked increase of the afterhyperpolarisation following each action potential is observed after prolonged sensory stimulation (Scuri et al., 2002). The afterhyperpolarisation is considered to have two major functions: it limits the firing frequency of the neuron and is responsible for generating spike-frequency adaptation. An activation of Ih should alleviate both effects, and indeed, we observed a shorter afterhyperpolarisation after Ih activation (Fig. 11B) as well as the occasional spontaneous generation of one or several action potentials upon repolarisation. Hyperpolarisations activate Ih, allowing counteracting depolarisations, so that the Em is shifted back to its original value. Vice versa, depolarisations will cause deactivation of Ih, which would result in a counteracting hyperpolarisation. Additionally, after prolonged mechanosensory stimulation, and the subsequent strong hyperpolarisation, the activation of Ih could help to preserve the effect of mechanosensory inputs by increasing the excitability. Although this could be a common trait of mechanosensory neurons in the leech, it is unclear whether this is a necessity for somatosensory signalling (i.e. a mechanism to reduce spike-frequency adaptation) or an amplifying mechanism for post-stimulus intervals (providing increased sensitivity to stimuli). The idea of shunting inhibition by reducing the input resistance of P neurons is refuted somewhat by the fact that deactivation of Ih is slow: before it is complete, e.g. during depolarisation from a hyperpolarised Em, Ih increases the neurons' excitability by a further depolarising tail potential. Additionally, it could be argued that increased excitability and shortened afterhyperpolarisation increase the neurons' firing rate, thereby possibly providing a more reliable sensory relay.

Although Ih is probably active near rest, the blockade with Cs+ did not hyperpolarise Em. This could be explained by an unspecific action of Cs+. K+ channels that may be active in leech P neurons could be blocked, e.g. the delayed rectifier and inward rectifier K+ channels (Pape, 1996). This would mask the hyperpolarisation caused by blocking Ih with a depolarisation caused by blocking K+ conductances. To what extent this occurs in leech P neurons is, however, unclear.

     
  • AE

    annulus erector (neuron)

  •  
  • AP

    anterior pagoda (neuron)

  •  
  • cAMP

    cyclic adenosine monophosphate

  •  
  • EK

    equilibrium potential for K+

  •  
  • Em

    resting membrane potential

  •  
  • ENa

    equilibrium potential for Na+

  •  
  • Erev

    reversal potential

  •  
  • gNa/gK

    conductance ratio

  •  
  • HN

    heart interneuron

  •  
  • IBMX

    3-isobutyl-1-methylxanthine

  •  
  • Ih

    hyperpolarisation-activated cation current

  •  
  • Ii

    instantaneous current

  •  
  • ISS

    steady-state or slow inward current

  •  
  • N

    noxious (neuron)

  •  
  • P

    pressure (neuron)

  •  
  • pNa/pK

    permeability ratio

  •  
  • R

    universal gas constant

  •  
  • SLS

    standard leech saline

  •  
  • T

    touch (neuron)

  •  
  • T

    absolute temperature

  •  
  • z

    number of elementary charges transferred

  •  
  • ZD7288

    4-ethylphenylamino-1,2-dimethyl-6-methylaminopyrimidinium chloride

FUNDING

This research received no specific grant from any funding agency in the public, commercial, or not-for-profit sectors.

The authors thank Claudia Roderigo and Simone Durry for excellent technical assistance and Prof. Dr Thomas Budde and Dr Susan Sangha for helpful comments on the manuscript.

Aidley
D. J.
(
1989
).
The Physiology of Excitable Cells.
Cambridge
:
Cambridge University Press
.
Akasu
T.
,
Shoji
S.
,
Hasuo
H.
(
1993
).
Inward rectifier and low-threshold calcium currents contribute to the spontaneous firing mechanism in neurons of the rat suprachiasmatic nucleus
.
Pflügers Arch.
425
,
109
-
116
.
Angstadt
J. D.
(
1999
).
Persistent inward currents in cultured Retzius cells of the medicinal leech
.
J. Comp. Physiol. A
184
,
49
-
61
.
Angstadt
J. D.
,
Calabrese
R. L.
(
1989
).
A hyperpolarization-activated inward current in heart interneurons of the medicinal leech
.
J. Neurosci.
9
,
2846
-
2857
.
Arbas
E. A.
,
Calabrese
R. L.
(
1987a
).
Ionic conductances underlying the activity of interneurons that control heartbeat in the medicinal leech
.
J. Neurosci.
7
,
3945
-
3952
.
Arbas
E. A.
,
Calabrese
R. L.
(
1987b
).
Slow oscillations of membrane potential in interneurons that control heartbeat in the medicinal leech
.
J. Neurosci.
7
,
3953
-
3960
.
Arbas
E. A.
,
Calabrese
R. L.
(
1990
).
Leydig neuron activity modulates heartbeat in the medicinal leech
.
J. Comp. Physiol. A
167
,
665
-
671
.
Baylor
D. A.
,
Nicholls
J. G.
(
1969
).
Chemical and electrical synaptic connexions between cutaneous mechanoreceptor neurones in the central nervous system of the leech
.
J. Physiol.
203
,
591
-
609
.
Biel
M.
,
Wahl-Schott
C.
,
Michalakis
S.
,
Zong
X.
(
2009
).
Hyperpolarization-activated cation channels: from genes to function
.
Physiol. Rev.
89
,
847
-
885
.
Bobker
D. H.
,
Williams
J. T.
(
1989
).
Serotonin augments the cationic current Ih in central neurons
.
Neuron
2
,
1535
-
1540
.
Bolívar
J. J.
,
Tapia
D.
,
Arenas
G.
,
Castanon-Arreola
M.
,
Torres
H.
,
Galarraga
E.
(
2008
).
A hyperpolarization-activated, cyclic nucleotide-gated (Ih-like) cationic current and HCN gene expression in renal inner medullary collecting duct cells
.
Am. J. Physiol.
294
,
C893
-
C906
.
Budde
T.
,
White
J. A.
,
Kay
A. R.
(
1994
).
Hyperpolarization-activated Na+-K+ current (Ih) in neocortical neurons is blocked by external proteolysis and internal TEA
.
J. Neurophysiol.
72
,
2737
-
2742
.
Budde
T.
,
Coulon
P.
,
Pawlowski
M.
,
Japes
A.
,
Meuth
P.
,
Meuth
S. G.
,
Pape
H. C.
(
2008
).
Reciprocal modulation of Ih and ITASK in thalamocortical relay neurons by halothane
.
Pflügers Arch.
456
,
1061
-
1073
.
Coulon
P.
(
2005
).
Elektrophysiologie der Volumenregulation von Retzius-Neuronen im Zentralnervensystem des medizinischen Blutegels
.
PhD thesis
,
Heinrich-Heine-Universität
,
Düsseldorf, Germany
.
Coulon
P.
,
Wüsten
H. J.
,
Hochstrate
P.
,
Dierkes
P. W.
(
2008
).
Swelling-activated chloride channels in leech Retzius neurons
.
J. Exp. Biol.
211
,
630
-
641
.
DiFrancesco
D.
(
1982
).
Block and activation of the pace-maker channel in calf Purkinje fibres: effects of potassium, caesium and rubidium
.
J. Physiol.
329
,
485
-
507
.
Doan
T. N.
,
Kunze
D. L.
(
1999
).
Contribution of the hyperpolarization-activated current to the resting membrane potential of rat nodose sensory neurons
.
J. Physiol.
514
,
125
-
138
.
Gasparini
S.
,
DiFrancesco
D.
(
1997
).
Action of the hyperpolarization-activated current (Ih) blocker ZD 7288 in hippocampal CA1 neurons
.
Pflügers Arch.
435
,
99
-
106
.
Hestrin
S.
(
1987
).
The properties and function of inward rectification in rod photoreceptors of the tiger salamander
.
J. Physiol.
390
,
319
-
333
.
Hille
B.
(
2001
).
Ion Channels of Excitable Membranes.
Sunderland, MA
:
Sinauer Associates
.
Hintz
K.
(
1999
).
Elektrophysiologische charakterisierung der Mg2+-regulation bei identifizierten neuronen und neuropil-gliazellen im zentralnervensystem des medizinischen blutegels
.
PhD thesis
,
Heinrich-Heine-Universität
,
Düsseldorf, Germany
.
Ho
W. K.
,
Brown
H. F.
,
Noble
D.
(
1994
).
High selectivity of the i(f) channel to Na+ and K+ in rabbit isolated sinoatrial node cells
.
Pflügers Arch.
426
,
68
-
74
.
Hodgkin
A. L.
,
Huxley
A. F.
(
1952
).
A quantitative description of membrane current and its application to conduction and excitation of nerve
.
J. Physiol.
117
,
500
-
544
.
Jansen
J. K.
,
Nicholls
J. G.
(
1973
).
Conductance changes, an electrogenic pump and the hyperpolarization of leech neurones following impulses
.
J. Physiol.
229
,
635
-
655
.
Johansen
J.
,
Kleinhaus
A. L.
(
1986
).
Differential sensitivity of tetrodotoxin of nociceptive neurons in 4 species of leeches
.
J. Neurosci.
6
,
3499
-
3504
.
Kaupp
U. B.
,
Seifert
R.
(
2001
).
Molecular diversity of pacemaker ion channels
.
Annu. Rev. Physiol.
63
,
235
-
257
.
Klees
G.
,
Hochstrate
P.
,
Dierkes
P. W.
(
2005
).
Sodium-dependent potassium channels in leech P neurons
.
J. Membr. Biol.
208
,
27
-
38
.
Lucht
M.
(
1998
).
Elektrophysiologische und pharmakologische Charakterisierung von 5-Hydroxytryptamin-Rezeptoren und Second-messenger-Kaskaden bei identifizierten Neuronen des Blutegel-Nervensystems
.
PhD thesis
,
Heinrich-Heine-Universtität
,
Düsseldorf, Germany
.
Ludwig
A.
,
Budde
T.
,
Stieber
J.
,
Moosmang
S.
,
Wahl
C.
,
Holthoff
K.
,
Langebartels
A.
,
Wotjak
C.
,
Munsch
T.
,
Zong
X.
, et al. 
. (
2003
).
Absence epilepsy and sinus dysrhythmia in mice lacking the pacemaker channel HCN2
.
EMBO J.
22
,
216
-
224
.
Lupica
C. R.
,
Bell
J. A.
,
Hoffman
A. F.
,
Watson
P. L.
(
2001
).
Contribution of the hyperpolarization-activated current (Ih) to membrane potential and GABA release in hippocampal interneurons
.
J. Neurophysiol.
86
,
261
-
268
.
Macri
V.
,
Proenza
C.
,
Agranovich
E.
,
Angoli
D.
,
Accili
E. A.
(
2002
).
Separable gating mechanisms in a mammalian pacemaker channel
.
J. Biol. Chem.
277
,
35939
-
35946
.
Masino
M. A.
,
Calabrese
R. L.
(
2002
).
Period differences between segmental oscillators produce intersegmental phase differences in the leech heartbeat timing network
.
J. Neurophysiol.
87
,
1603
-
1615
.
McCormick
D. A.
,
Pape
H. C.
(
1990a
).
Properties of a hyperpolarization-activated cation current and its role in rhythmic oscillation in thalamic relay neurones
.
J. Physiol.
431
,
291
-
318
.
McCormick
D. A.
,
Pape
H. C.
(
1990b
).
Noradrenergic and serotonergic modulation of a hyperpolarization-activated cation current in thalamic relay neurones
.
J. Physiol.
431
,
319
-
342
.
Meuth
S. G.
,
Kanyshkova
T.
,
Meuth
P.
,
Landgraf
P.
,
Munsch
T.
,
Ludwig
A.
,
Hofmann
F.
,
Pape
H. C.
,
Budde
T.
(
2006
).
Membrane resting potential of thalamocortical relay neurons is shaped by the interaction among TASK3 and HCN2 channels
.
J. Neurophysiol.
96
,
1517
-
1529
.
Mistrík
P.
,
Pfeifer
A.
,
Biel
M.
(
2006
).
The enhancement of HCN channel instantaneous current facilitated by slow deactivation is regulated by intracellular chloride concentration
.
Pflügers Arch.
452
,
718
-
727
.
Nicholls
J.
(
1987
).
The search for connections
. In
Studies of Regeneration in the Nervous System of a Leech
,
Magnes Lecture Series
, Vol.
2
, p.
25
.
Sunderland, MA
:
Sinauer Associates
.
Pape
H. C.
(
1996
).
Queer current and pacemaker: the hyperpolarization-activated cation current in neurons
.
Annu. Rev. Physiol.
58
,
299
-
327
.
Pape
H. C.
,
McCormick
D. A.
(
1989
).
Noradrenaline and serotonin selectively modulate thalamic burst firing by enhancing a hyperpolarization-activated cation current
.
Nature
340
,
715
-
718
.
Pirtle
T. J.
,
Willingham
K.
,
Satterlie
R. A.
(
2010
).
A hyperpolarization-activated inward current alters swim frequency of the pteropod mollusk Clione limacina
.
Comp. Biochem. Physiol. A Mol. Integr. Physiol.
157
,
319
-
327
.
Proenza
C.
,
Angoli
D.
,
Agranovich
E.
,
Macri
V.
,
Accili
E. A.
(
2002
).
Pacemaker channels produce an instantaneous current
.
J. Biol. Chem.
277
,
5101
-
5109
.
Santoro
B.
,
Liu
D. T.
,
Yao
H.
,
Bartsch
D.
,
Kandel
E. R.
,
Siegelbaum
S. A.
,
Tibbs
G. R.
(
1998
).
Identification of a gene encoding a hyperpolarization-activated pacemaker channel of brain
.
Cell
93
,
717
-
729
.
Schlue
W. R.
(
1991
).
Effects of ouabain on intracellular ion activities of sensory neurons of the leech central nervous system
.
J. Neurophysiol.
65
,
736
-
746
.
Schlue
W. R.
,
Deitmer
J. W.
(
1980
).
Extracellular potassium in neuropile and nerve cell body region of the leech central nervous system
.
J. Exp Biol.
87
,
23
-
43
.
Schlue
W. R.
,
Deitmer
J. W.
(
1984
).
Potassium distribution and membrane potential of sensory neurons in the leech nervous system
.
J. Neurophysiol.
51
,
689
-
704
.
Scuri
R.
,
Mozzachiodi
R.
,
Brunelli
M.
(
2002
).
Activity-dependent increase of the AHP amplitude in T sensory neurons of the leech
.
J. Neurophysiol.
88
,
2490
-
2500
.
Shin
K. S.
,
Rothberg
B. S.
,
Yellen
G.
(
2001
).
Blocker state dependence and trapping in hyperpolarization-activated cation channels: evidence for an intracellular activation gate
.
J. Gen. Physiol.
117
,
91
-
101
.
Siddall
M. E.
,
Trontelj
P.
,
Utevsky
S. Y.
,
Nkamany
M.
,
Macdonald
K. S.
(
2007
).
Diverse molecular data demonstrate that commercially available medicinal leeches are not Hirudo medicinalis
.
Proc. R. Soc. Lond B
274
,
1481
-
1487
.
Tobin
A. E.
,
Calabrese
R. L.
(
2005
).
Myomodulin increases Ih and inhibits the Na/K pump to modulate bursting in leech heart interneurons
.
J. Neurophysiol.
94
,
3938
-
3950
.
Wollmuth
L. P.
,
Hille
B.
(
1992
).
Ionic selectivity of Ih channels of rod photoreceptors in tiger salamanders
.
J. Gen. Physiol.
100
,
749
-
765
.
Wuttke
W. A.
,
Berry
M. S.
(
1992
).
Modulation of inwardly rectifying Na+-K+ channels by serotonin and cyclic nucleotides in salivary gland cells of the leech, Haementeria
.
J. Membr. Biol.
127
,
57
-
68
.