Next Article in Journal
Structural and Spectroscopic Properties of Isoconazole and Bifonazole—Experimental and Theoretical Studies
Next Article in Special Issue
Characterization of a Neutral Sphingomyelinase Activity in Human Serum and Plasma
Previous Article in Journal
Asymmetric Synthesis of Trifluoroethyl-Based, Chiral 3-Benzyloxy-1- and -2-Phenyl-quinazolinones of Biomedicinal Interest by Radical Type Cross-Coupling to Olefins
Previous Article in Special Issue
Counteracting Immunosuppression in the Tumor Microenvironment by Oncolytic Newcastle Disease Virus and Cellular Immunotherapy
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sedative Properties of Dexmedetomidine Are Mediated Independently from Native Thalamic Hyperpolarization-Activated Cyclic Nucleotide-Gated Channel Function at Clinically Relevant Concentrations

Department of Anesthesiology and Intensive Care Medicine, Klinikum Rechts der Isar, Technical University of Munich School of Medicine, 81675 Munich, Germany
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(1), 519; https://doi.org/10.3390/ijms24010519
Submission received: 24 November 2022 / Revised: 19 December 2022 / Accepted: 20 December 2022 / Published: 28 December 2022
(This article belongs to the Special Issue State-of-the-Art Biochemistry in Germany)

Abstract

:
Dexmedetomidine is a selective α2-adrenoceptor agonist and appears to disinhibit endogenous sleep-promoting pathways, as well as to attenuate noradrenergic excitation. Recent evidence suggests that dexmedetomidine might also directly inhibit hyperpolarization-activated cyclic-nucleotide gated (HCN) channels. We analyzed the effects of dexmedetomidine on native HCN channel function in thalamocortical relay neurons of the ventrobasal complex of the thalamus from mice, performing whole-cell patch-clamp recordings. Over a clinically relevant range of concentrations (1–10 µM), the effects of dexmedetomidine were modest. At a concentration of 10 µM, dexmedetomidine significantly reduced maximal Ih amplitude (relative reduction: 0.86 [0.78–0.91], n = 10, and p = 0.021), yet changes to the half-maximal activation potential V1/2 occurred exclusively in the presence of the very high concentration of 100 µM (−4,7 [−7.5–−4.0] mV, n = 10, and p = 0.009). Coincidentally, only the very high concentration of 100 µM induced a significant deceleration of the fast component of the HCN activation time course (τfast: +135.1 [+64.7–+151.3] ms, n = 10, and p = 0.002). With the exception of significantly increasing the membrane input resistance (starting at 10 µM), dexmedetomidine did not affect biophysical membrane properties and HCN channel-mediated parameters of neuronal excitability. Hence, the sedative qualities of dexmedetomidine and its effect on the thalamocortical network are not decisively shaped by direct inhibition of HCN channel function.

1. Introduction

Dexmedetomidine (DEX) is commonly used for short- and longer-term sedation of adult patients receiving treatment in intensive care units and as an anesthetic adjunct due to its generally favorable pharmacological profile, including anxiolytic, analgesic, and opioid-saving properties [1,2]. When examining the effects of DEX on neural network activity, it appears as though it induces a rather unique brain state. Electroencephalographic (EEG) recordings of patients being administered DEX are characterized by slow-wave (SW, 0.1–1 Hz) and delta (δ, 1–4 Hz) oscillations, as well as spindle activity (12–16 Hz) [3]. Despite perfunctory similarities with EEG features observable during general anesthesia with propofol, which induces profound unconsciousness, DEX shifts brain activity in a rather different, sleep-like state [4]. Patients who receive DEX for sedation, can be aroused readily, maintain the ability to communicate and cooperate properly, and further do not display some of the adverse effects generally accompanied by other hypnotic agents, such as respiratory depression [2,5].Among a long list of potential suspects, a relatively small number of key molecular targets of general anesthetics have been identified [6]. Deviating from this canonical compilation, DEX predominantly acts as a selective agonist of presynaptic alpha-2 (α2)-adrenergic receptors, which are located on axonal projections of glutamatergic neurons originating in the locus coeruleus (LC). By decreasing noradrenaline release in a dose-dependent manner [7], DEX results in hyperpolarization and reduced excitability of LC neurons [8]. Inhibition of LC neurons is thought to partly entail the promotion of endogenous non-rapid eye movement sleep (NREM)-promoting pathways [9] with manifold consequences on a network level [5], including the sleep-promoting disinhibition of the hypothalamic ventrolateral preoptic nucleus (VLPO), which in turn inhibits the tuberomammillary nucleus as part of the ascending monoaminergic arousal center via γ-aminobutyric acid (GABA)-ergic and galanin-ergic VLPO projections [9,10,11,12]. Of the three distinct subtypes of α2-adrenergic receptors identified (α2A, α2B, and α2C), α2A-receptor subtypes appear to primarily mediate sedative, hypnotic, as well as anesthesia-sparing effects in mice [13,14,15].
Further effects of DEX include attenuated excitation of arousal-related basal forebrain regions, the thalamus and cortex [16], as well as antinociception both at the spinal cord level as well as at supraspinal sites [17]. The application of sedative concentrations of DEX results in reduced thalamocortical (TC) functional connectivity, a diminished cerebral metabolic rate of glucose, as well as an attenuated cerebral blood flow, preferentially in the thalamus [18].
As the LC heavily projects both onto the thalamus and the cortex [19], the observed changes are likely due to a DEX-mediated dampening of ascending arousal pathways, resulting in hyperpolarization of TC relay neurons and thereby interrupting TC communication [20]. Hyperpolarization of TC relay neurons is an important requisite for generating and maintaining synchronized δ-oscillations within the TC network, as these neurons transition from a tonic, single-spike activity to a rhythmic, burst-firing mode [6,21]. In addition to hyperpolarization as a result of dampened ascending input, DEX might also directly modulate the intrinsic excitability of TC relay neurons. It is established that properties of thalamic oscillations and the pacemaker function of TC relay neurons are decisively shaped by the inwardly-directed, mixed cation current Ih [22,23], which is mediated by hyperpolarization-activated cyclic nucleotide-gated channels (HCN).
How HCN channels mediated excitability of TC relay neurons is affected by anesthetic agents has been extensively investigated over the past decades [6,24] and a number of in vitro studies demonstrate inhibition of native HCN channel function in the cases of propofol [25], pentobarbital [26], xenon [27], halothane [28,29], and sevoflurane [30]. Interestingly, recent evidence suggests that DEX might also act on heterologously expressed HCN channels and attenuate HCN channel function in neurons from dorsal root ganglia [31,32]. This is further supported by the notion that other selective α2-adrenoceptor agonists, including clonidine, have been shown to inhibit HCN channel function [33,34,35,36]. DEX-mediated effects on the neurophysiology of the brain might, therefore, arise from direct action on VB neurons, in addition to the effects that are elicited by the modulation of neurons originating for the LC, primarily via α2A-receptor subtypes [14].
Hence, we investigated whether DEX, at clinically relevant concentrations, affects native HCN channel function and HCN channel-mediated excitability of TC relay neurons located in the primarily first-order ventrobasal (VB) complex of the thalamus [37], using the patch-clamp technique in acutely prepared brain slices from mice.

2. Results

2.1. Effects of DEX on Membrane Properties of TC Neurons

Passive and active membrane biophysics were derived from current–voltage relationship recordings. Under baseline recordings, TC neurons displayed a median resting membrane potential of −58.2 (IQR: −59.7–−56.9) mV, n = 31. For all three DEX concentration groups, we did not detect a significant change in the resting membrane potential after 30 min and 45 min of DEX application, see Supplementary Table S1. The effects of DEX are presented as an absolute difference from the baseline recordings (Δ 1 µM DEX: −0.31 [−1.5–+1.2] mV, n = 11, p > 0.999; Δ 10 µM DEX: +0.8 [−1.9–+1.5] mV, n = 10, p > 0.999; and Δ 100 µM DEX: −0.8 [−2.6–+0.2] mV, n = 10, p > 0.342). In between concentration groups, there were no significant differences with regard to baseline conditions (p = 0.671, Kruskal–Wallis test with a posthoc Dunn’s test). A total of 40 µM of HCN channel blocker ZD7288, which has a reported half maximal inhibitory concentration (IC50) of 15 µM [38], induced a strong hyperpolarization of the resting membrane potential in all cells measured, which persisted after the subsequent application of 10 µM DEX (control: −60.9 [−63.2–−58.5] mV]; +40 µM ZD7288: −65.5 [−74.0–−61.0] mV; and +10 µm DEX: −63.5 [−72.1–−61.0] mV, n = 5). See Figure 1A.
Beginning at a concentration of 10 µM DEX, we observed a significant increase in the membrane input resistance, indicative of decreased neuronal leakiness. Under control conditions, the median input resistance was at 343.4 (252.3–365.6) MΩ, n = 31. The effects of DEX are presented as an absolute difference from the baseline recordings (Δ 1 µM DEX: +24.0 [−10.0–+37.2] MΩ, n = 11, p = 0.602; Δ 10 µM DEX: +27.9 [+9.6–+89.0] MΩ, n = 10, p = 0.011; and Δ 100 µM DEX: +95.3 [+73.2–+120.5] MΩ, n = 10, p < 0.001). A total of 40 µM ZD7288 also resulted in an increase in the input resistance in all cells recorded, which again persisted after applying 10 µM DEX (an absolute difference from the baseline Δ +40 µM ZD7288: +140.3 [+63.9–+340.5] MΩ and +10 µm DEX: +159.1 [+34.0–+378.3] MΩ, n = 5) Figure 1B. However, when analyzing changes over the whole current–voltage relationship, the increase of the input resistance in presence of 10 µM DEX was only statistically significant during hyperpolarizing current injections, in the range from −90 pA to −70 pA (an absolute change of the input resistance after 10 µM DEX at −70 pA current injection compared to the baseline: +38.2 [+19.3–+53.4] MΩ, n = 10, p = 0.037 and at −60 pA: +40 [+3.2–+64.9] MΩ, n = 10, p = 0.084).
When analyzing the effect of DEX on current–voltage relationships of recorded cells, which were determined by injecting successively depolarizing currents (10 pA increments), only the high concentration of 100 µM DEX persistently induced a significant hyperpolarization in the whole range between −90 and +350 pA, whereas 10 µM DEX only induced significant changes during strongly hyperpolarizing currents, see Figure 1C. Current–voltage relationships were not recorded in the presence of agents, which selectively prevent the occurrence of action potentials. This might be reflected by broader interquartile ranges of voltage responses obtained by injection of rather high current pulses.
Similarly, intrinsic firing properties of TC neurons were only significantly affected by a very high concentration of 100 µM DEX. Under the baseline conditions, the threshold potential for action potential firing was determined at −39.2 (−40.5–−37.6) mV. 100 µM DEX shifted this threshold toward depolarization, with no significant effects occurring at lower concentrations (Δ 1 µM DEX: −0.6 [−2.6–−0.2] mV, n = 11, p = 0.583; Δ 10 µM DEX: −0.4 [−2.4–+0.9] mV, n = 10, p = 0.850; and Δ 100 µM DEX: +1.3 [+1.1–+5.4] mV, n = 10, p = 0.013). In contrast, HCN channel blocker ZD7288 shifted the threshold potential toward hyperpolarization. The co-application of ZD7288 and 10 µM DEX did not reverse the observed hyperpolarization (control: −38.1 [−39.4–−36.7] mV; Δ 40 µM ZD7288: −4.8 [−7.6–−2.4] mV; and Δ 10 µM DEX + ZD7288: −3.7 [−7.3–−2.4] mV, n = 5), see Figure 1D.
We further analyzed the effect of DEX on the maximum frequency of tonic action potential firing. The frequency was derived from the current injection of +190 pA, as we did not observe a further increase in maximum frequency beyond this value. Median tonic action potential firing at the baseline conditions occurred at 40.2 (36.3–44.9) Hz, n = 31. Again, significant changes were exclusively observable in the presence of the high DEX concentration, with a decrease in tonic action potential frequency (Δ 1 µM DEX: +2.3 [−1.8–°+5.6] Hz n = 11, p = 0.953; Δ 10 µM DEX: −0.7 [−2.2–+3.1] Hz, n = 10, p > 0.999; and Δ 100 µM DEX: −12.8 [−18.7–−9.7] Hz, n = 10, p < 0.001), see Figure 1E.
All experiments were conducted in the presence of 150 µM Ba2+ to preclude interference with Kir and K2P channels. On native recordings, the addition of Ba2+ resulted in a small depolarization of the resting membrane potential (control: −57.6 [−59.5–−57.4] mV vs. Ba2+: −56.2 [−58.7–−55.8] mV, n = 7, p = 0.016) and an increase in the input resistance (control: 180.6 [133.7–197.6] MΩ vs. Ba2+: 229.9 [212.5–263.2] MΩ, n = 7, p = 0.016), whereas there were no significant changes regarding the action potential threshold (p = 0.578) as well as tonic action potential firing (p = 0.094).

2.2. Influence of DEX on HCN Channel Function and Ih Currents

We analyzed the effects of DEX on the HCN channel-mediated, mixed cation current Ih, as well as modalities of HCN channel function. Since the extent of HCN channel activation correlates with increasing hyperpolarization [39], the Ih current amplitude derived at −133 mV was considered the maximum (Ih max) current amplitude, and Ih current amplitudes evoked at more depolarized values were set in relation accordingly, see Figure 2A. Both 10 µM and 100 µM DEX applications resulted in a significant reduction in Ih max (values are relative to the normalized baseline recordings; 1 µM DEX: 0.86 [0.79–0.98], n = 11, p = 0.068; 10 µM DEX: 0.86 [0.78–0.91], n = 10, p = 0.021; and 100 µM DEX: 0.78 [0.50–0.84], n = 10, p < 0.001). By contrast, HCN channel blocker ZD7288 (40 µM) led to an almost complete reduction in Ih max in all cells recorded (values relative to the normalized baseline recordings; 40 µM ZD7288: 0.05 [0.02–0.06], n = 5), see Figure 2B. At a membrane potential of −73 mV, reflecting more physiological conditions with only partial HCN channel activation, only 100 µM DEX induced a significant decrease (1 µM DEX: 0.92 [0.86–1.05], n = 11, p > 0.999; 10 µM DEX: 0.90 [0.80–1.00], n = 10, p = 0.293; 100 µM DEX: 0.57 [0.42–0.58], n = 10, p < 0.001; and 40 µM ZD7288: 0.05 [0.03–0.14], n = 5), see Figure 2C. However, in the case of 10 µM DEX, the reduction in Ih current amplitude became significant beyond membrane potentials of −83 mV.
Furthermore, we evaluated whether DEX affects the voltage-dependent activation of HCN channels and open channel probability via a tail current analysis. The respective Itail activation curves fitted with a Boltzmann function are shown in Figure 3A. We observed a median half-maximal activation potential (V1/2) of −86.6 (−85.2–−88.0) mV for TC relay neurons under baseline conditions. Only in the presence of the high concentration of 100 µM DEX was the V1/2 shifted toward hyperpolarization in accordance with a reduced open-channel probability (Δ 1 µM DEX: +0.3 [−0.9–+1.1] mV, n = 11, p > 0.999; Δ 10 µM DEX: +0.4 [−1.5–+1.7] mV, n = 10, p > 0.999; and Δ 100 µM DEX: -4.7 [−7.5–−4.0] mV, n = 10, p = 0.009), see Figure 3B.
In comparison to native recordings, the addition of 150 µM Ba2+ did not significantly change Ih max (control: 976.7 [552.6–1291.4] pA vs. Ba2+: 916.0 [516.8–1128.4] pA, n = 6, p = 0.844) nor V1/2 (control: −87.1 [−90.0–−83.4] mV vs. Ba2+: −86.0 [−88.6–−84.9] pA, n = 6, p = 0.844).
A similar pattern of only modest effects at high DEX concentrations was seen with regard to the time course of Ih activation, which was evaluated by fitting the Ih current recorded at −133 mV to a biexponential function, see Figure 3C. Absolute values of the corresponding activation time constants τfast and τslow change considerably, depending on the voltage step analyzed, as well as the duration of the current recording. Due to the limited effects observed with regard to Ih amplitude, we chose to apply the biexponential fitting to the maximum voltage step at −133 mV. The median value of τfast at −133 mV was 231 ms (214–273 ms; n = 31), while τslow was 1143 ms (986–1405 ms; n = 31) under control conditions. Only the fast component of HCN activation kinetics was subjected to a significant deceleration, while the slow component remained, overall, unaffected. Likewise, the observed deceleration was only significant at a high concentration of 100 µM DEX (τfast: Δ 1 µM DEX: −4.1 [−20.3–+4.9] ms, n = 11, p > 0.999; Δ 10 µM DEX: +25.5 [+4.1–+39.6] ms, n = 10, p = 0.232; Δ 100 µM DEX: +135.1 [+64.7–+151.3], n = 10, p = 0.002; τslow: Δ 1 µM DEX: +93.0 [−205.9–−198.4] ms, n = 11, p > 0.999; Δ 10 µM DEX: +163.8 [−150.4–+574.2] ms, n = 10, p > 0.999; and Δ 100 µM DEX: −208.2 [−697.5–+155.0], n = 10, p = 0.382), see Figure 3C.

2.3. Effects of DEX on TC Neuron Firing Properties

Lastly, we evaluated whether DEX affects modalities of neuronal excitability in TC relay neurons generally associated with HCN channel function. Therefore, we performed additional whole-cell current-clamp recordings. Upon injection of a hyperpolarizing current (−350 pA, 500 ms), TC neurons display an anomalous rectification of the membrane potential, termed “voltage sag”, followed by a low-threshold, calcium burst of action potential firing. Upon burst cessation, the neurons show an afterdepolarization. Both afterdepolarization and voltage sag are largely mediated by HCN channels [40], see Figure 4A. All current-clamp recordings were recorded in the presence of 150 µM Ba2+. In a subset of experiments, we evaluated the impact of Ba2+ on native current-clamp recordings. As shown in Figure 4A, Ba2+ unmasks an otherwise discreet voltage sag (control: 14.8 [5.7–18.7] mV vs. Ba2+: 58.9 [53.2–63.9], n = 7, p = 0.016) and significantly accelerates the median rebound burst delay (control: 62.5 [49.0–86.0] ms vs. Ba2+: 33.5 [29.5–44.0] ms, n = 7, p = 0.016). In comparison, HCN channel blocker ZD7288 completely attenuates the inwardly directed voltage rectification. Further, in two out of five cells, we were not able to induce bursting activity in the presence of ZD7288, even when increasing the current injection to −500 pA or compensating for the ZD7288-induced membrane hyperpolarization. In comparison, the effects of DEX were negligible, see Figure 4B.
The median voltage sag amplitude displayed by TC neurons under baseline conditions was 86.4 (70.7–92.9) mV (n = 31). Interestingly, even the high concentration of 100 µM DEX, which was sufficient to elicit a significant change in Ih amplitude, half-maximal activation potential, as well as a deceleration of the fast activation kinetics time constant, did not affect the voltage sag (Δ 1 µM DEX: −0.4 [−2.9–+1.3] mV, n = 11, p > 0.999; Δ 10 µM DEX: −1.0 [−4.8–+4.4] mV, n = 10, p = 0.753; and Δ 100 µM DEX: +5.1 [+0.4–+12.4], n = 10, p = 0.083), see Figure 4C.
Significant changes with regard to rebound delay, the number of action potentials during rebound spike activation, as well as median burst duration exclusively occurred at the high concentration of 100 µM DEX. During control conditions, the median delay of rebound burst-spiking upon hyperpolarization was 29.0 ms (27.0–30.8 ms; n = 31), with a significant prolongation of 6.5 ms (5.1–8.9) ms (Δ 100 µM DEX: n = 10, p = 0.003; lower concentrations not shown), see Figure 4D. The median number of action potentials observed during rebound burst-firing was 9.5 (7.3–12.5) under control conditions, with a significant reduction of −6.8 (−8.8–−4.7) action potentials (Δ 100 µM DEX: n = 10, p = 0.002, see Figure 4E). Corresponding with the lower number of action potentials, the median duration of rebound bursts (under control conditions of 225.5 [126.5–399.0] ms) was only significantly reduced at the high concentration (Δ 100 µM DEX: −140.3 [−247.1–−72.8] ms, n = 10, p = 0.016), see Figure 4F.
In conclusion, our data demonstrate that DEX does not significantly decrease voltage sag amplitudes or attenuate rebound burst-firing in TC relay neurons, which are thought to be governed by an HCN conductance, at least not at clinically relevant concentrations.
Absolute values of all parameters analyzed in this study before and after DEX application are presented in Supplementary Tables S2–S13 for each concentration group separately. The supplementary tables further include corresponding paired statistical analysis.

3. Discussion

We used patch-clamp recordings of TC relay neurons from the VB thalamic complex to investigate whether DEX can directly inhibit Ih or HCN channel-mediated parameters of neuronal excitability ex vivo. Our principal finding is that the effects of DEX in clinically relevant concentrations on native HCN channel function are modest.
A total of 10 µM DEX significantly reduced Ih max and increased the neuronal input resistance, albeit only in the range of hyperpolarizing current injections. HCN channels decisively regulate the resting membrane potential of VB TC relay neurons [41] and, therefore, inhibitory effects are to be expected primarily in the hyperpolarized membrane range. However, we did not observe hyperpolarization of the resting membrane potential and there were no significant changes in the half-maximal activation voltage within the clinically relevant concentration range. The relationship between pharmacological HCN channel inhibition, neuronal input resistance, and resting membrane potential is not always straightforward. Xenon inhibited HCN channels and increased neuronal input resistance, but did not hyperpolarize the resting membrane potential of TC relay neurons [27]. Propofol and sevoflurane both inhibit HCN channels and hyperpolarize the resting membrane potential, but concomitantly decrease neuronal input resistance, likely via GABAA-receptor activation [25,30,42]. With the inclusion of barium, we minimized the influence of Kir and subfamilies of K2P channels [40], but we cannot positively preclude the activation of further conductances. Furthermore, the reduction in Ih observed at −133 mV at 10 µM DEX might not be physiologically relevant enough to translate into changes in the resting membrane potential, especially as changes of Ih at −73 mV were not statistically significant. This is further supported by the observation that DEX did not attenuate the inwardly rectifying voltage sag, in contrast to other anesthetics [25,27,30]. As the resting membrane potential remained unaffected, all current-clamp recordings were performed at the resting membrane potential rather than different fixed voltages. In comparison, HCN channel blocker ZD7288 hyperpolarized the resting membrane potential and increased the neuronal input resistance of all cells recorded, while simultaneously diminishing Ih over the whole voltage range. Furthermore, our findings with regard to the very high concentration of 100 µM DEX should be interpreted with caution, as we cannot positively exclude unspecific receptor interactions or the occurrence of other conductances.
An obvious concern with respect to in vitro studies relates to the question of whether the applied concentrations of a pharmacological agent are indeed clinically relevant. A recent study [43] states that the clinically relevant DEX concentration range lies from 1 to 2.5 µM in the case of rats. A total of 5 µM was already regarded as the upper end of the scale. Incidentally noted, this study also demonstrated that DEX appears to be non-cytotoxic and did not induce cell death in cultured neonatal rat neurons. Consistent herewith, the half-maximal effective concentration (EC50) of DEX required for sedation and analgesia in rats was determined to be at 1 µM [44]. Some sources assume considerably lower dosages (0.001–1 µM DEX) to be relevant [45]. However, mice appear to generally require a higher minimum DEX dose to induce sedation [9]. Therefore, in vitro studies investigating the effect of DEX in mice neurons tend to apply slightly higher concentrations. In mice dorsal root ganglion neurons, DEX reduced capsaicin-induced intracellular calcium concentrations in a dose-dependent manner between 1 and 50 µM DEX [46]. In vitro studies explicitly investigating the effect of DEX on HCN channel function have also applied mostly similar concentrations (0.1–10 µM DEX [31,32], 0.1–1 µM DEX [47], and 10 nM–6 µM DEX [48]). We, therefore, consider the applied concentrations of 1 and 10 µM DEX in this study to be clinically relevant with regard to a hypothesized effect on HCN channels, while 100 µM certainly represents a supratherapeutic concentration. It is also necessary to ensure that DEX can sufficiently penetrate the ex vivo brain slice preparation, especially when comparing our results to in vitro studies. Two studies using brain slice preparations mention that DEX, as well as the pharmacologically similar drug clonidine, reduced Ih amplitude already within 5 min after application [35,48]. Our experiments were conducted over the comparatively long timespan of 45 min (the results presented here reflect the comparison between baseline measurements and DEX application after 45 min) and any DEX-induced changes should have occurred within this interval. We further conducted recordings after 30 min, but there were no significant differences when compared to 45 min. Additionally, the previously reported DEX-induced inhibition of Ih and HCN channel function was characterized as non-recoverable even after a prolonged washout [48].
The goal of this study was to investigate whether DEX can directly inhibit HCN channel function. One indication of this assumption was based on the observation that clonidine, another α2-adrenoceptor agonist, has been shown to directly inhibit HCN2 and HCN4 (as well as HCN1, albeit with less sensitivity) channels in transfected HEK293 cells and block Ih in isolated sinoatrial node pacemaker cells. Comparable to the HCN blocking agent ZD7288, clonidine also shifted V1/2 toward more hyperpolarized potentials [34]. In dopaminergic neurons of the ventral tegmental area, clonidine decreased Ih amplitude and slowed its rate of activation independently from cAMP via the activation of protein kinase C [35]. Furthermore, clonidine has hitherto been shown to inhibit HCN-mediated Ih in further neuronal cell types, including rat trigeminal root ganglion neurons [49], rat hypoglossal motoneurons [33], and rat pyramidal neurons in the prefrontal cortex [50].
So far, our knowledge about the effects of DEX on HCN channels is still limited. There is indirect evidence from a study using an in vivo model of peripheral nerve blockade in rats [51]. Here, adding DEX to the local anesthetic ropivacaine prolonged analgesia compared to analgesia with ropivacaine alone. Prima facie, our findings are contrasted by the results of several electrophysiological in vitro studies. One group investigated the effects of DEX on HCN channel function in view of neuropathic pain and analgesia, complementary to behavioral tests [31,32]. The authors report that DEX, in concentrations ranging from 0.1 to 10 µM, inhibits both mHCN1 and mHCN2 channels expressed in HEK293 cells. Significant differences between HCN1 and HCN2 in relative DEX inhibition were not found [31]. In a follow-up study, the same authors report patch-clamp recordings from HEK293 cells expressing HCN1 and HCN2 [32]. Another study [47] analyzed the effect of DEX on cultured rat pituitary tumor cells (GH3). DEX induced a concentration-dependent reduction in the amplitude of Ih (IC50 of 1.21 µM) and a substantial hyperpolarizing shift in the voltage-dependent activation course of Ih (−10 mV). Further, DEX attenuated Ih in pheochromocytoma PC12 cells and slowed its activation time course. Endocrine and neuroendocrine cells, including GH3 cells, intrinsically express isoforms HCN2 and HCN3 [52]. Lastly, a study investigated the effect of DEX on hypothalamic paraventricular nucleus (PVN) neurons from rat brain slice preparations [48]. DEX reduced the Ih amplitude irreversibly and the V1/2 was shifted toward hyperpolarization by −8 mV This effect was only marginally dependent on intracellular cAMP. Rat PVN neurons have been shown to express mRNA of HCN1, HCN2, and HCN3, but not HCN4 [53].
Furthermore, DEX dose-dependently hyperpolarized PVN neurons and decreased input resistance, but this hyperpolarization was blocked by pretreatment with 0.3 mM Ba2+ [48]. Therefore, the hyperpolarizing effect of DEX is likely facilitated by Kir and certain potassium K2P channels [54]. This could explain the absence of hyperpolarization in our recordings, as all our experiments were conducted in presence of Ba2+. In contrast with the reported decrease in input resistance in the study by Shirasaka et al. [48], we observed a significant increase in membrane input resistance beginning at 10 µM. This increase in input resistance would be, in principle, consistent with HCN channel inhibition [39]. However, the observed increase in the neuronal input resistance at 10 µM DEX was only significant during strongly hyperpolarizing current injections. We previously reported a similar effect in the case of xenon [27]. Many anesthetic agents, such as propofol and sevoflurane, decrease neuronal input resistance, probably via concomitant allosteric activation of GABAA-receptors [25,30,55]. This assumption is further supported by the observation that xenon both inhibits Ih and increases input resistance [27] while eliciting only minimal GABAergic responses [56]. However, propofol, sevoflurane, and xenon clearly attenuated native HCN channel function in native TC relay neurons.
As a matter of principle, recordings of HCN-mediated Ih are known to be strongly sensitive to changes in experimental conditions, including the ionic composition of extra- and intra-cellular solutions, temperature, pH, and patch-clamp configuration [39]. In contrast to heterologously expressed HCN channel isoforms or cultured cells, cells from ex vivo brain slice preparations are further integrated into complex modulatory networks. Accordingly, ex vivo preparations have been an integral and well-appreciated component of anesthesia research over the last decades [57]. However, we previously demonstrated pharmacological inhibition of HCN channel function in TC neurons, both in the cases of xenon and sevoflurane, using the same method and brain slice preparation [27,30]. Instead, we would like to argue that an important aspect of the question, of why our findings diverge from aforementioned studies [31,32,47,48], might relate to HCN channel subunit composition. The HCN family encompasses four different homologous subunits (HCN1–HCN4). These subunits assemble as homo- and hetero-tetramers [58] and differ with respect to their pharmacological profile and biophysical properties [58,59,60]. These quantitative differences include the degree of cAMP regulation, voltage-dependence of activation, and steady-state properties, as well as activation time constants [39]. HCN channels and corresponding Ih currents are present throughout the mammalian nervous system, but the distribution of specific HCN isoforms and their combinations display a more heterogeneous pattern [39]. TC relay neurons of the VB complex predominantly express HCN2 and HCN4 isoforms [61,62] and, correspondingly, studies demonstrated that a proper Ih conductance in TC relay neurons is largely dependent on HCN2 [63] and also HCN4 [64]. Importantly, subunit selective modulation of HCN channels by anesthetic and other pharmacological agents has been demonstrated before [28,65,66]. With regard to their activation kinetics, HCN subunits act quite differently [39]. As the opening kinetics accelerate with increasing hyperpolarization [59], specific τ values depend on the hyperpolarization step they were derived from. HCN2 displays intermediate opening kinetics compared to other isoforms and its τ-value ranges from 150 ms to 1 s (−140 mV to −70 mV), whereas the pronouncedly slower HCN4 isoform ranges from several hundred milliseconds to seconds [59,66]. As a restrictive factor, it should be noted that a prolonged membrane hyperpolarization of −110 mV or beyond endangers the vitality of neurons in acute brain slice preparations [67]. As in previous studies [27,30], we used a protocol with variable voltage-step durations, given the fact that HCN channel activation accelerates with increasing hyperpolarization [59]. However, our recordings might underestimate some effects, as we potentially did not reach a complete steady-state equilibrium during strongly hyperpolarizing voltage steps.
V1/2 values for HCN2 and HCN4 range between −70 and −100 mV [58] and, in the case of both isoforms, their steady-state activation is readily modulated by cAMP. The intracellular addition of cAMP can shift the V1/2 10 to 25 mV toward depolarization [68]. In this study, TC neurons of the VB complex displayed a median V1/2 of −86.6 mV, as well as a median τslow of 1143 ms. Together with the existing research evidence, our data strongly indicate that the recorded conductance stems from HCN2 and HCN4, whereas the other studies cited include recordings of HCN1 and HCN2, HCN1, HCN2 and HCN3, and HCN2 andHCN3. It is further noteworthy that the DEX-induced inhibition of Ih in slice recordings of rat hypothalamic PVN neurons was almost completely antagonized by the application of yohimbine, an α2-adrenoceptor antagonist. This suggests that the inhibition of Ih was reliant on an intact α2-adrenoceptor mechanism [48].
The non-inactivating, inwardly rectifying current Ih plays a crucial role in the generation, amplification, and stable maintenance of rhythmic oscillations within the TC network [69] and further controls the resting membrane potential of TC relay neurons [41]. The intrinsic properties of thalamic HCN channels and their finely tuned interplay with other conductances, such as the low-threshold Ca2+ current IT [21], promote the generation of oscillations in the delta frequency band, both in individual TC relay neurons [70], and within the TC network [71]. EEG recordings of patients receiving DEX sedation are characterized by slow delta oscillations superimposed with, at lower concentrations, sleep spindles, and this state is further distinguished by a noteworthy absence of profound unconsciousness [3,4]. Interestingly, these molecular actions converge on a preferential disruption of TC functional connectivity, while corticocortical functional connectivity was largely maintained [18]. Both EEG hallmarks of DEX sedation, slow delta oscillations and dose-dependent occurrence of spindle activity, are thought to originate within TC loop circuits [5]. This is further supported by the observation that DEX administration sufficient for a sedative state induces high power in the δ frequency band in local field potentials from thalamic nuclei. However, compared with the first-order VB complex, the higher-order central medial thalamus seems to be preferentially affected [20]. As DEX administration results in sedation and patients remain cooperative or can be easily aroused [72], TC interaction, especially communication between thalamic first-order nuclei of the VB complex (which are the focus of this study) and the primary somatosensory cortex, must preserve a certain degree of intactness. If DEX would indeed be able to strongly inhibit HCN channel function in TC relay neurons, with subsequent changes of biophysical membrane properties, then one would rather expect a profound disruption of sensory information relaying. Moreover, it has been demonstrated that DEX-induced sedation and hypnosis in mice are primarily mediated by the α2A-adrenergic receptor subtype [13,14,73]. Most pharmacological and physiological functions of α2-adrenergic receptors are ascribed to this specific subtype [15]. Interestingly, thalamic nuclei appear to primarily (albeit lightly) express the α2B-adrenergic receptor subtype [74]. Therefore, our observation that DEX does not potently inhibit thalamic HCN channel function is consistent with previous studies demonstrating the importance of α2A-adrenergic receptors in DEX-mediated sedation.
Naturally, the findings of this study should be interpreted in light of its limitations. Firstly, all recordings in this study were performed at room temperature. Albeit not uncommon [27,28,30,31,32,47,48], there is evidence that HCN channel activation and deactivation time constants significantly accelerate with increasing temperatures, whereas HCN channel-mediated current amplitudes were either not significantly changed by the temperature or the effect of temperature adjustments was much lower [75,76]. Further, all animals used in this study were female. There is limited evidence that in certain neuronal cell populations, HCN-mediated Ih might be regulated by circulating estradiol in a cyclical manner [77], though this does not appear to be a general concern so far [60,67]. However, sex-specific differences in HCN channel modulation should be more adequately addressed in future studies.
Ultimately, our data demonstrate that DEX, in a clinically relevant concentration range, does not significantly inhibit the native HCN channel function of TC neurons in the thalamic VB complex ex vivo, and parameters of neuronal excitability remain largely unaffected by its application. In comparison with anesthetic agents such as propofol, xenon, halothane, and sevoflurane, which substantially inhibit native HCN channel function in TC relay neurons in vitro [25,27,29,30], DEX administration does not as potently disrupt thalamic neuronal activity. Consistent with our patch-clamp findings, DEX-mediated slow delta oscillations and spindles exhibit smaller amplitudes compared with propofol [4]. These differences on a molecular level, including a different effect on native HCN channel function, could be part of the explanation for why DEX administration results in a sedation-like state, while other anesthetic agents can induce profound unconsciousness.

4. Materials and Methods

4.1. TC Slice Preparation

Animal handling was reviewed and approved by the competent veterinary office (Munich, Germany). We used female C57Bl/6N mice (P28–P35), which were put in deep isoflurane anesthesia (as indicated by loss of righting reflex) before surgical brain extraction. Brain slices with a 350 µm-thickness were prepared according to a well-established protocol [78], which allows for reliable identification of the VB complex of the thalamus, using a vibratome (HM 650 V, Thermo Fisher Scientific, Walldorf, Germany). Upon removal, brains were rapidly transferred into the ice-cold artificial cerebrospinal fluid (aCSF) optimized for brain slice preparation, containing (in mM) 125.0 NaCl, 2.5 KCl, 1.25 NaH2PO4, 25.0 D-glucose, 25.0 NaHCO3, 6.0 MgCl2, and 0.5 CaCl2. During the cutting procedure, aCSF was continuously aerated with carbogen (95% O2, 5% CO2) to ensure oxygenation and a stable pH of approximately 7.4. Slices containing the region of interest (3–4 per brain) were then transferred into a storage chamber containing a carbogen-saturated standard aCSF (in mM: 125.0 NaCl, 2.5 KCl, 1.25 NaH2PO4, 25.0 D-glucose, 25.0 NaHCO3, 1.0 MgCl2, and 2.0 CaCl2) and incubated for a minimum of 30 min at 34 °C in a warm water bath, followed by an additional 30 min of recovery time at room temperature.

4.2. Experimental Setup and DEX Application

Individual slices were transferred into the recording chamber of the patch-clamp setup and mechanically fixated with a custom-made grid. During experiments, slices were continuously perfused with carbogen-saturated standard aCSF at a flow rate of 5–8 mL/min. Upon visual allocation of the thalamic VB (Zeiss, Oberkochen, Germany), individual TC relay neurons were identified with infrared videomicroscopy (Hamamatsu Photonics, Herrsching, Germany). We established whole-cell recordings of single cells with fire-polished, borosilicate glass pipettes (Sutter Instrument, Novato, CA, USA). Pipettes were fabricated with open tip resistances between 4 and 6 MΩ, using a Flaming/Brown micropipette puller (P-1000, Sutter Instrument, Novato, CA, USA) and filled with an intracellular solution with the following composition (in mM): 130.0 K-D-gluconate, 5.0 NaCl, 2.0 MgCl2, 10.0 HEPES, 0.5 EGTA, 2.0 K2-ATP, 0.3 Na2-GTP, and pH: 7.25. As an electrophysiological defining feature, TC relay neurons exhibit an Ih-dependent inward rectification of the membrane potential following injection of hyperpolarizing currents, termed voltage sag [40], which we used for the identification of the targeted neurons. After ensuring stable baseline recording conditions for at least 15 min, DEX was added to the aerated reservoir to attain three effective concentrations: 1 µM, 10 µM, and 100 µM. For the preparation of stock solutions, DEX hydrochloride was purchased from Sigma-Aldrich (Steinheim, Germany) and dissolved in aqua dest. for stock concentrations of 1 mM and 10 mM, whereas for the 100 µM concentration, DEX was directly dissolved in the aCSF of the circulation reservoir for each experiment, due to the solubility limitations of this concentration. Every brain slice was exclusively exposed to one DEX concentration. In order to ensure sufficient pharmacological slice penetration, we allowed for 30 min of DEX exposure before repeating recording protocols. The second set of recordings was performed after an additional 15 min (45 min after the beginning of the exposure). For approximately half of the recordings, we further conducted a washout by circulating fresh carbogen-saturated aCSF through the recording chamber for about 30 min before repeating the recordings. All experiments were recorded at room temperature (21–24 °C). All salts and chemicals used in these experiments were acquired from Sigma-Aldrich (Steinheim, Germany), with the exception of ZD7288 (Tocris, Wiesbaden-Nordenstadt, Germany).

4.3. Electrophysiology

All patch-clamp recordings were conducted in the whole-cell configuration with a discontinuous single-electrode voltage-clamp (dSEVC) amplifier (SEC 10L, NPI Electronic, Tamm, Germany). Though series resistance is effectively compensated by applying the dSEVC technique with high switching frequencies (60–80 kHz and a 25% duty cycle), it was continuously monitored. Ih currents were recorded in the voltage-clamp mode, according to a protocol previously described [27,30]. Starting from a membrane potential of −43 mV, Ih currents of TC neurons were gradually activated by applying hyperpolarizing voltage steps (10 voltage steps of 10 mV each, up to a maximum of −133 mV). As the stability of whole-cell recordings tends to deteriorate at highly hyperpolarized membrane potentials and factoring in the concurrent acceleration of Ih activation kinetics through hyperpolarization, we opted for incrementally reducing the pulse length of each voltage step by 500 ms, resulting in a 2.0 s pulse duration at −133 mV [29,41]. Ih appears to reliably reach its maximum at membrane potentials beyond −110 mV [79]. Therefore, maximal Ih current amplitude was derived from the −133 mV voltage step and used for normalization. Each voltage step also included a tail step of −103 mV before returning to the baseline of −43 mV. Respective tail current amplitudes (Itail) were normalized to the tail current acquired from the −133 mV voltage step and were used to assess the voltage dependency of steady-state activation. For this purpose, the equation (I–Imin)/(Imax–Imin) was applied, with Imax representing the tail current amplitude derived from the voltage step of −133 mV to −103 mV and Imin derived from the voltage step from −43 mV to −103 mV. These data fit well with a Boltzmann distribution, which is usually applied to estimate steady-state activation. The voltage-dependent time course of Ih activation was approximated by fitting a biexponential function to the currents evoked during the hyperpolarizing step of −133 mV, yielding fast (τfast) and slow (τslow) time constants of Ih activation, as previously described for TC relay neurons [80]. Fitting with two exponentials is known to improve fits for a majority of Ih currents evoked at strongly hyperpolarized membrane potentials, while a monoexponential fit approach yields more robust results at more depolarized membrane potentials [81]. As we did not analyze time course activation as a function of membrane potential, we opted for applying a biexponential fit.
Further current-clamp recordings were recorded from the same TC neurons. After injection of a hyperpolarizing current of −350 pA for a duration of 500 ms at the resting membrane potential, TC neurons display an inwardly directed rectification of the membrane potential, termed voltage sag. We defined the voltage sag as the difference between the maximum voltage deflection and the steady-state potential. Subsequently, we measured the time between the start of the membrane repolarization at the end of the hyperpolarizing current injection and the peak of the first action potential during rebound burst-firing (rebound burst delay). This current-clamp protocol was repeated eight times for each recording and later analyzed as a graphically averaged composite trace with Igor 5 (WaveMetrics, Lake Oswego, OR, USA). The impact of DEX on active and passive electrical membrane properties, comprising resting membrane potential, membrane input resistance, the current–voltage relationship, the threshold value of action potential firing, and the tonic action potential frequency were derived from membrane voltage changes induced by successive increases of current injections (starting from −90 pA to +360 pA, 10 pA increase per injection with a 500 ms duration) at the resting membrane potential. For current-clamp recordings, we only adjusted the resting membrane to control (pre-drug application) levels when significant changes in the resting membrane potential were observed.
The liquid junction potential was corrected online, as estimated based on the ionic compositions of intra- and extra-cellular solutions and with help of the Liquid Junction Potential Calculator tool (Clampex 10.7, Molecular Devices. San Jose, CA, USA). As previously elaborated on ([27,30], all experiments were recorded in presence of 0.15 mM barium chloride (Ba2+), which was added to the aCSF circulation. Briefly summarized, Ba2+ inhibits the potentially interfering effects on inwardly rectifying potassium channels (Kir) and members of the two-pore domain acid-sensitive potassium channels (K2P), as well as facilitates improved analysis by significantly enhancing voltage sag amplitude [40].
The current responses were amplified, low-pass filtered at 3 kHz, digitized (ITC-16 Computer Interface, Instrutech Corporation, Great Neck, NY, USA) with a sampling frequency of 9 kHz, and stored on a hard drive (HP ProDesk 400 G2.5 SFF, Hewlett-Packard, Palo Alto, CA, USA). For data acquisition, we used Pulse version 8.5 (HEKA Elektronik, Lambrecht, Germany). Data from individual cells were only included in the analysis if they did not meet one of the following exclusion criteria: (1) no GΩ-seal resistance formation upon establishment of cell-attached configuration, (2) an initial resting membrane potential below −50 mV, (3) an unstable series resistance or holding current (>20% change), and (4) no stable baseline conditions after 15 min, as assessed by the experimenter. Analysis of electrophysiological recordings was performed offline with Igor 5 (WaveMetrics, Lake Oswego, OR, USA). Action potentials were evaluated using Mini Analysis (Synaptosoft, Fort Lee, NJ, USA).

4.4. Statistical Analysis

For statistical evaluation of the data obtained and the creation of graphs, we used GraphPad Prism 7.03 (GraphPad Software, San Diego, CA, USA). Sample sizes for different concentration groups were chosen based on previous experience. As formally testing for normal distribution inherently poses a challenge in case of limited sample sizes [82], we opted for the application of non-parametrical testing by default. Accordingly, all results are presented as medians with the interquartile range (IQR). We used the Kruskal–Wallis test for independent groups with Dunn’s posthoc test to correct for multiple comparisons. In the first step, we tested for significant differences between the medians of the control values of the three different concentration groups to ensure comparability between groups. In the second step, we tested for significant differences between absolute changes of the three concentration groups compared to a normalized control. For multiple comparisons of repeated measures, we used the non-parametric Friedman test. In the case of simple comparisons, we used the Wilcoxon matched-pairs signed-rank test. For sample sizes with n ≤ 5, we only provide a descriptive statistic, as non-parametric statistical tests cannot be meaningfully applied in these cases. Where applicable, p-values were adjusted for multiplicity and differences were considered statistically significant if p < 0.05, as graphically indicated by asterisks.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ijms24010519/s1.

Author Contributions

Conceptualization, R.H. and S.K.; methodology, S.S., S.K. and R.H.; validation, R.H. and S.K.; formal analysis, S.S., C.W. and M.K.; investigation, S.S., C.W. and C.K.; resources, G.S., R.H. and S.K.; data curation, M.K. and R.H.; writing—original draft preparation, S.S., C.W. and S.K.; writing—review and editing, S.S., G.S., M.K., S.K. and R.H.; visualization, S.S., C.W., C.K. and M.K. supervision, R.H. and S.K.; project administration, R.H. and S.K. All authors have read and agreed to the published version of the manuscript.

Funding

This work was exclusively supported by institutional and departmental funding from the Technical University of Munich.

Institutional Review Board Statement

Care and handling of the animals were reviewed and approved by the competent veterinary office (Munich, Germany). As animals were exclusively used for scientific organ extractions (brains), this study did not include animal experiments by definition of the German law (section 7, subsection 2, and sentence 3 Tierschutzgesetz).

Informed Consent Statement

Not applicable.

Data Availability Statement

The raw data supporting the conclusions of this article will be made available by the corresponding author upon reasonable request, without undue reservation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Gerresheim, G.; Schwemmer, U. Dexmedetomidin. Anaesthesist 2013, 62, 661–674. [Google Scholar] [CrossRef] [PubMed]
  2. Keating, G.M. Dexmedetomidine: A Review of Its Use for Sedation in the Intensive Care Setting. Drugs 2015, 75, 1119–1130. [Google Scholar] [CrossRef] [PubMed]
  3. Huupponen, E.; Maksimow, A.; Lapinlampi, P.; Särkelä, M.; Saastamoinen, A.; Snapir, A.; Scheinin, H.; Scheinin, M.; Meriläinen, P.; Himanen, S.-L.; et al. Electroencephalogram spindle activity during dexmedetomidine sedation and physiological sleep. Acta Anaesthesiol. Scand. 2008, 52, 289–294. [Google Scholar] [CrossRef]
  4. Akeju, O.; Pavone, K.J.; Westover, M.B.; Vazquez, R.; Prerau, M.J.; Harrell, P.G.; Hartnack, K.E.; Rhee, J.; Sampson, A.L.; Habeeb, K.; et al. A comparison of propofol- and dexmedetomidine-induced electroencephalogram dynamics using spectral and coherence analysis. Anesthesiology 2014, 121, 978–989. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Purdon, P.L.; Sampson, A.; Pavone, K.J.; Brown, E.N. Clinical Electroencephalography for Anesthesiologists: Part I: Background and Basic Signatures. Anesthesiology 2015, 123, 937–960. [Google Scholar] [CrossRef] [Green Version]
  6. Franks, N.P. General anaesthesia: From molecular targets to neuronal pathways of sleep and arousal. Nat. Rev. Neurosci. 2008, 9, 370–386. [Google Scholar] [CrossRef]
  7. Jorm, C.M.; Stamford, J.A. Actions of the hypnotic anaesthetic, dexmedetomidine, on noradrenaline release and cell firing in rat locus coeruleus slices. Br. J. Anaesth. 1993, 71, 447–449. [Google Scholar] [CrossRef]
  8. Nacif-Coelho, C.; Correa-Sales, C.; Chang, L.L.; Maze, M. Perturbation of ion channel conductance alters the hypnotic response to the alpha 2-adrenergic agonist dexmedetomidine in the locus coeruleus of the rat. Anesthesiology 1994, 81, 1527–1534. [Google Scholar] [CrossRef]
  9. Nelson, L.E.; Lu, J.; Guo, T.; Saper, C.B.; Franks, N.P.; Maze, M. The alpha2-adrenoceptor agonist dexmedetomidine converges on an endogenous sleep-promoting pathway to exert its sedative effects. Anesthesiology 2003, 98, 428–436. [Google Scholar] [CrossRef] [Green Version]
  10. Sherin, J.E.; Elmquist, J.K.; Torrealba, F.; Saper, C.B. Innervation of histaminergic tuberomammillary neurons by GABAergic and galaninergic neurons in the ventrolateral preoptic nucleus of the rat. J. Neurosci. 1998, 18, 4705–4721. [Google Scholar] [CrossRef]
  11. Morairty, S.; Rainnie, D.; McCarley, R.; Greene, R. Disinhibition of ventrolateral preoptic area sleep-active neurons by adenosine: A new mechanism for sleep promotion. Neuroscience 2004, 123, 451–457. [Google Scholar] [CrossRef] [PubMed]
  12. Saper, C.B.; Scammell, T.E.; Lu, J. Hypothalamic regulation of sleep and circadian rhythms. Nature 2005, 437, 1257–1263. [Google Scholar] [CrossRef] [PubMed]
  13. Lakhlani, P.P.; MacMillan, L.B.; Guo, T.Z.; McCool, B.A.; Lovinger, D.M.; Maze, M.; Limbird, L.E. Substitution of a mutant alpha2a-adrenergic receptor via "hit and run" gene targeting reveals the role of this subtype in sedative, analgesic, and anesthetic-sparing responses in vivo. Proc. Natl. Acad. Sci. USA 1997, 94, 9950–9955. [Google Scholar] [CrossRef] [Green Version]
  14. Philipp, M.; Brede, M.; Hein, L. Physiological significance of alpha(2)-adrenergic receptor subtype diversity: One receptor is not enough. Am. J. Physiol. Regul. Integr. Comp. Physiol. 2002, 283, R287-95. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Gilsbach, R.; Hein, L. Are the pharmacology and physiology of α2 adrenoceptors determined by α2-heteroreceptors and autoreceptors respectively? Br. J. Pharmacol. 2012, 165, 90–102. [Google Scholar] [CrossRef] [PubMed]
  16. España, R.A.; Berridge, C.W. Organization of noradrenergic efferents to arousal-related basal forebrain structures. J. Comp. Neurol. 2006, 496, 668–683. [Google Scholar] [CrossRef] [PubMed]
  17. Panzer, O.; Moitra, V.; Sladen, R.N. Pharmacology of sedative-analgesic agents: Dexmedetomidine, remifentanil, ketamine, volatile anesthetics, and the role of peripheral Mu antagonists. Anesthesiol. Clin. 2011, 29, 587–605. [Google Scholar] [CrossRef]
  18. Akeju, O.; Loggia, M.L.; Catana, C.; Pavone, K.J.; Vazquez, R.; Rhee, J.; Contreras Ramirez, V.; Chonde, D.B.; Izquierdo-Garcia, D.; Arabasz, G.; et al. Disruption of thalamic functional connectivity is a neural correlate of dexmedetomidine-induced unconsciousness. eLife 2014, 3, e04499. [Google Scholar] [CrossRef] [Green Version]
  19. Jones, B.E.; Moore, R.Y. Ascending projections of the locus coeruleus in the rat. II. Autoradiographic study. Brain Res. 1977, 127, 25–53. [Google Scholar] [CrossRef]
  20. Baker, R.; Gent, T.C.; Yang, Q.; Parker, S.; Vyssotski, A.L.; Wisden, W.; Brickley, S.G.; Franks, N.P. Altered activity in the central medial thalamus precedes changes in the neocortex during transitions into both sleep and propofol anesthesia. J. Neurosci. 2014, 34, 13326–13335. [Google Scholar] [CrossRef]
  21. McCormick, D.A.; Bal, T. Sleep and arousal: Thalamocortical mechanisms. Annu. Rev. Neurosci. 1997, 20, 185–215. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Yue, B. The role of H-current in regulating strength and frequency of thalamic network oscillations. Thalamus Relat. Syst. 2001, 1, 95–103. [Google Scholar] [CrossRef] [PubMed]
  23. Destexhe, A.; Sejnowski, T.J. Interactions between membrane conductances underlying thalamocortical slow-wave oscillations. Physiol. Rev. 2003, 83, 1401–1453. [Google Scholar] [CrossRef] [PubMed]
  24. Goldstein, P.A. HCN1 Channels as Targets for Volatile Anesthetics: Coming to the Fore. Anesth. Analg. 2015, 121, 594–596. [Google Scholar] [CrossRef] [PubMed]
  25. Ying, S.-W.; Abbas, S.Y.; Harrison, N.L.; Goldstein, P.A. Propofol block of I(h) contributes to the suppression of neuronal excitability and rhythmic burst firing in thalamocortical neurons. Eur. J. Neurosci. 2006, 23, 465–480. [Google Scholar] [CrossRef] [PubMed]
  26. Wan, X.; Mathers, D.A.; Puil, E. Pentobarbital modulates intrinsic and GABA-receptor conductances in thalamocortical inhibition. Neuroscience 2003, 121, 947–958. [Google Scholar] [CrossRef] [PubMed]
  27. Mattusch, C.; Kratzer, S.; Buerge, M.; Kreuzer, M.; Engel, T.; Kopp, C.; Biel, M.; Hammelmann, V.; Ying, S.-W.; Goldstein, P.A.; et al. Impact of Hyperpolarization-activated, Cyclic Nucleotide-gated Cation Channel Type 2 for the Xenon-mediated Anesthetic Effect: Evidence from In Vitro and In Vivo Experiments. Anesthesiology 2015, 122, 1047–1059. [Google Scholar] [CrossRef] [Green Version]
  28. Chen, X.; Sirois, J.E.; Lei, Q.; Talley, E.M.; Lynch, C.; Bayliss, D.A. HCN subunit-specific and cAMP-modulated effects of anesthetics on neuronal pacemaker currents. J. Neurosci. 2005, 25, 5803–5814. [Google Scholar] [CrossRef] [Green Version]
  29. Budde, T.; Coulon, P.; Pawlowski, M.; Meuth, P.; Kanyshkova, T.; Japes, A.; Meuth, S.G.; Pape, H.-C. Reciprocal modulation of I (h) and I (TASK) in thalamocortical relay neurons by halothane. Pflugers Arch. 2008, 456, 1061–1073. [Google Scholar] [CrossRef]
  30. Schwerin, S.; Kopp, C.; Pircher, E.; Schneider, G.; Kreuzer, M.; Haseneder, R.; Kratzer, S. Attenuation of Native Hyperpolarization-Activated, Cyclic Nucleotide-Gated Channel Function by the Volatile Anesthetic Sevoflurane in Mouse Thalamocortical Relay Neurons. Front. Cell. Neurosci. 2020, 14, 606687. [Google Scholar] [CrossRef]
  31. Yang, Y.; Meng, Q.; Pan, X.; Xia, Z.; Chen, X. Dexmedetomidine produced analgesic effect via inhibition of HCN currents. Eur. J. Pharmacol. 2014, 740, 560–564. [Google Scholar] [CrossRef] [PubMed]
  32. Yang, Y.; Xia, Z.; Meng, Q.; Liu, K.; Xiao, Y.; Shi, L. Dexmedetomidine relieves neuropathic pain by inhibiting hyperpolarization-activated cyclic nucleotide-gated currents in dorsal root ganglia neurons. Neuroreport 2018, 29, 1001–1006. [Google Scholar] [CrossRef] [PubMed]
  33. Parkis, M.A.; Berger, A.J. Clonidine reduces hyperpolarization-activated inward current (Ih) in rat hypoglossal motoneurons. Brain Res. 1997, 769, 108–118. [Google Scholar] [CrossRef] [PubMed]
  34. Knaus, A.; Zong, X.; Beetz, N.; Jahns, R.; Lohse, M.J.; Biel, M.; Hein, L. Direct inhibition of cardiac hyperpolarization-activated cyclic nucleotide-gated pacemaker channels by clonidine. Circulation 2007, 115, 872–880. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Inyushin, M.U.; Arencibia-Albite, F.; Vázquez-Torres, R.; Vélez-Hernández, M.E.; Jiménez-Rivera, C.A. Alpha-2 noradrenergic receptor activation inhibits the hyperpolarization-activated cation current (Ih) in neurons of the ventral tegmental area. Neuroscience 2010, 167, 287–297. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Won, J.; Lee, P.R.; Oh, S.B. Alpha 2 adrenoceptor agonist guanabenz directly inhibits hyperpolarization-activated, cyclic nucleotide-modulated (HCN) channels in mesencephalic trigeminal nucleus neurons. Eur. J. Pharmacol. 2019, 854, 320–327. [Google Scholar] [CrossRef] [PubMed]
  37. Sherman, S.M.; Guillery, R.W. Functional organization of thalamocortical relays. J. Neurophysiol. 1996, 76, 1367–1395. [Google Scholar] [CrossRef] [Green Version]
  38. Wu, X.; Liao, L.; Liu, X.; Luo, F.; Yang, T.; Li, C. Is ZD7288 a selective blocker of hyperpolarization-activated cyclic nucleotide-gated channel currents? Channels 2012, 6, 438–442. [Google Scholar] [CrossRef] [Green Version]
  39. Biel, M.; Wahl-Schott, C.; Michalakis, S.; Zong, X. Hyperpolarization-activated cation channels: From genes to function. Physiol. Rev. 2009, 89, 847–885. [Google Scholar] [CrossRef] [Green Version]
  40. Datunashvili, M.; Chaudhary, R.; Zobeiri, M.; Lüttjohann, A.; Mergia, E.; Baumann, A.; Balfanz, S.; Budde, B.; van Luijtelaar, G.; Pape, H.-C.; et al. Modulation of Hyperpolarization-Activated Inward Current and Thalamic Activity Modes by Different Cyclic Nucleotides. Front. Cell. Neurosci. 2018, 12, 369. [Google Scholar] [CrossRef]
  41. Meuth, S.G.; Kanyshkova, T.; Meuth, P.; Landgraf, P.; Munsch, T.; Ludwig, A.; Hofmann, F.; Pape, H.-C.; Budde, T. Membrane resting potential of thalamocortical relay neurons is shaped by the interaction among TASK3 and HCN2 channels. J. Neurophysiol. 2006, 96, 1517–1529. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Chen, X.; Shu, S.; Bayliss, D.A. Suppression of ih contributes to propofol-induced inhibition of mouse cortical pyramidal neurons. J. Neurophysiol. 2005, 94, 3872–3883. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  43. Jimenez-Tellez, N.; Iqbal, F.; Pehar, M.; Casas-Ortiz, A.; Rice, T.; Syed, N.I. Dexmedetomidine does not compromise neuronal viability, synaptic connectivity, learning and memory in a rodent model. Sci. Rep. 2021, 11, 16153. [Google Scholar] [CrossRef] [PubMed]
  44. Huang, R.; Hertz, L. Receptor subtype and dose dependence of dexmedetomidine-induced accumulation of [14C]glutamine in astrocytes suggests glial involvement in its hypnotic-sedative and anesthetic-sparing effects. Brain Res. 2000, 873, 297–301. [Google Scholar] [CrossRef] [PubMed]
  45. Guo, F.; Ding, Y.; Yu, X.; Cai, X. Effect of dexmedetomidine, midazolam, and propofol on lipopolysaccharide-stimulated dendritic cells. Exp. Ther. Med. 2018, 15, 5487–5494. [Google Scholar] [CrossRef] [Green Version]
  46. Lee, B.-M.; Jang, Y.; Park, G.; Kim, K.; Oh, S.H.; Shin, T.J.; Chung, G. Dexmedetomidine modulates transient receptor potential vanilloid subtype 1. Biochem. Biophys. Res. Commun. 2020, 522, 832–837. [Google Scholar] [CrossRef]
  47. Lu, T.-L.; Lu, T.-J.; Wu, S.-N. Effectiveness in Block by Dexmedetomidine of Hyperpolarization-Activated Cation Current, Independent of Its Agonistic Effect on α2-Adrenergic Receptors. Int. J. Mol. Sci. 2020, 21, 9110. [Google Scholar] [CrossRef]
  48. Shirasaka, T.; Kannan, H.; Takasaki, M. Activation of a G protein-coupled inwardly rectifying K+ current and suppression of Ih contribute to dexmedetomidine-induced inhibition of rat hypothalamic paraventricular nucleus neurons. Anesthesiology 2007, 107, 605–615. [Google Scholar] [CrossRef] [Green Version]
  49. Takeda, M.; Ikeda, M.; Tanimoto, T.; Lipski, J.; Matsumoto, S. Changes of the excitability of rat trigeminal root ganglion neurons evoked by α2-adrenoreceptors. Neuroscience 2002, 115, 731–741. [Google Scholar] [CrossRef]
  50. Carr, D.B.; Andrews, G.D.; Glen, W.B.; Lavin, A. alpha2-Noradrenergic receptors activation enhances excitability and synaptic integration in rat prefrontal cortex pyramidal neurons via inhibition of HCN currents. J. Physiol. 2007, 584, 437–450. [Google Scholar] [CrossRef]
  51. Brummett, C.M.; Hong, E.K.; Janda, A.M.; Amodeo, F.S.; Lydic, R. Perineural dexmedetomidine added to ropivacaine for sciatic nerve block in rats prolongs the duration of analgesia by blocking the hyperpolarization-activated cation current. Anesthesiology 2011, 115, 836–843. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Stojilkovic, S.S.; Tabak, J.; Bertram, R. Ion channels and signaling in the pituitary gland. Endocr. Rev. 2010, 31, 845–915. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  53. Qiu, D.-L.; Chu, C.-P.; Shirasaka, T.; Tsukino, H.; Nakao, H.; Kato, K.; Kunitake, T.; Katoh, T.; Kannan, H. Corticotrophin-releasing factor augments the I(H) in rat hypothalamic paraventricular nucleus parvocellular neurons in vitro. J. Neurophysiol. 2005, 94, 226–234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Päsler, D.; Gabriel, S.; Heinemann, U. Two-pore-domain potassium channels contribute to neuronal potassium release and glial potassium buffering in the rat hippocampus. Brain Res. 2007, 1173, 14–26. [Google Scholar] [CrossRef]
  55. Ying, S.W.; Goldstein, P.A. Propofol-block of SK channels in reticular thalamic neurons enhances GABAergic inhibition in relay neurons. J Neurophysiol 2005, 93, 1935–1948. [Google Scholar] [CrossRef] [Green Version]
  56. Kubota, H.; Akaike, H.; Okamitsu, N.; Jang, I.-S.; Nonaka, K.; Kotani, N.; Akaike, N. Xenon modulates the GABA and glutamate responses at genuine synaptic levels in rat spinal neurons. Brain Res. Bull. 2020, 157, 51–60. [Google Scholar] [CrossRef]
  57. Voss, L.J.; García, P.S.; Hentschke, H.; Banks, M.I. Understanding the Effects of General Anesthetics on Cortical Network Activity Using Ex Vivo Preparations. Anesthesiology 2019, 130, 1049–1063. [Google Scholar] [CrossRef] [Green Version]
  58. Altomare, C.; Terragni, B.; Brioschi, C.; Milanesi, R.; Pagliuca, C.; Viscomi, C.; Moroni, A.; Baruscotti, M.; DiFrancesco, D. Heteromeric HCN1-HCN4 channels: A comparison with native pacemaker channels from the rabbit sinoatrial node. J. Physiol. 2003, 549, 347–359. [Google Scholar] [CrossRef]
  59. Ludwig, A.; Zong, X.; Stieber, J.; Hullin, R.; Hofmann, F.; Biel, M. Two pacemaker channels from human heart with profoundly different activation kinetics. EMBO J 1999, 18, 2323–2329. [Google Scholar] [CrossRef] [Green Version]
  60. He, C.; Chen, F.; Li, B.; Hu, Z. Neurophysiology of HCN channels: From cellular functions to multiple regulations. Prog. Neurobiol. 2014, 112, 1–23. [Google Scholar] [CrossRef]
  61. Postea, O.; Biel, M. Exploring HCN channels as novel drug targets. Nat. Rev. Drug Discov. 2011, 10, 903–914. [Google Scholar] [CrossRef] [PubMed]
  62. Notomi, T.; Shigemoto, R. Immunohistochemical localization of Ih channel subunits, HCN1-4, in the rat brain. J. Comp. Neurol. 2004, 471, 241–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Ludwig, A.; Budde, T.; Stieber, J.; Moosmang, S.; Wahl, C.; Holthoff, K.; Langebartels, A.; Wotjak, C.; Munsch, T.; Zong, X.; et al. Absence epilepsy and sinus dysrhythmia in mice lacking the pacemaker channel HCN2. EMBO J. 2003, 22, 216–224. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Zobeiri, M.; Chaudhary, R.; Blaich, A.; Rottmann, M.; Herrmann, S.; Meuth, P.; Bista, P.; Kanyshkova, T.; Lüttjohann, A.; Narayanan, V.; et al. The Hyperpolarization-Activated HCN4 Channel is Important for Proper Maintenance of Oscillatory Activity in the Thalamocortical System. Cereb. Cortex 2019, 29, 2291–2304. [Google Scholar] [CrossRef] [Green Version]
  65. Chen, X.; Shu, S.; Kennedy, D.P.; Willcox, S.C.; Bayliss, D.A. Subunit-specific effects of isoflurane on neuronal Ih in HCN1 knockout mice. J. Neurophysiol. 2009, 101, 129–140. [Google Scholar] [CrossRef] [Green Version]
  66. Novella Romanelli, M.; Sartiani, L.; Masi, A.; Mannaioni, G.; Manetti, D.; Mugelli, A.; Cerbai, E. HCN Channels Modulators: The Need for Selectivity. Curr. Top. Med. Chem. 2016, 16, 1764–1791. [Google Scholar] [CrossRef] [Green Version]
  67. Riegelhaupt, P.M.; Tibbs, G.R.; Goldstein, P.A. HCN and K2P Channels in Anesthetic Mechanisms Research; Elsevier: Amsterdam, The Netherlands, 2018. [Google Scholar]
  68. Wainger, B.J.; DeGennaro, M.; Santoro, B.; Siegelbaum, S.A.; Tibbs, G.R. Molecular mechanism of cAMP modulation of HCN pacemaker channels. Nature 2001, 411, 805–810. [Google Scholar] [CrossRef]
  69. Amarillo, Y.; Mato, G.; Nadal, M.S. Analysis of the role of the low threshold currents IT and Ih in intrinsic delta oscillations of thalamocortical neurons. Front. Comput. Neurosci. 2015, 9, 52. [Google Scholar] [CrossRef] [Green Version]
  70. Leresche, N.; Jassik-Gerschenfeld, D.; Haby, M.; Soltesz, I.; Crunelli, V. Pacemaker-like and other types of spontaneous membrane potential oscillations of thalamocortical cells. Neurosci. Lett. 1990, 113, 72–77. [Google Scholar] [CrossRef]
  71. Steriade, M.; Timofeev, I. Neuronal Plasticity in Thalamocortical Networks during Sleep and Waking Oscillations. Neuron 2003, 37, 563–576. [Google Scholar] [CrossRef]
  72. Kasuya, Y.; Govinda, R.; Rauch, S.; Mascha, E.J.; Sessler, D.I.; Turan, A. The correlation between bispectral index and observational sedation scale in volunteers sedated with dexmedetomidine and propofol. Anesth. Analg. 2009, 109, 1811–1815. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  73. Hunter, J.C.; Fontana, D.J.; Hedley, L.R.; Jasper, J.R.; Lewis, R.; Link, R.E.; Secchi, R.; Sutton, J.; Eglen, R.M. Assessment of the role of alpha2-adrenoceptor subtypes in the antinociceptive, sedative and hypothermic action of dexmedetomidine in transgenic mice. Br. J. Pharmacol. 1997, 122, 1339–1344. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Scheinin, M.; Lomasney, J.W.; Hayden-Hixson, D.M.; Schambra, U.B.; Caron, M.G.; Lefkowitz, R.J.; Fremeau, R.T. Distribution of α2-adrenergic receptor subtype gene expression in rat brain. Brain Res. Mol. Brain Res. 1994, 21, 133–149. [Google Scholar] [CrossRef] [PubMed]
  75. Magee, J.C. Dendritic Hyperpolarization-Activated Currents Modify the Integrative Properties of Hippocampal CA1 Pyramidal Neurons. J. Neurosci. 1998, 18, 7613–7624. [Google Scholar] [CrossRef] [Green Version]
  76. Rateau, Y.; Ropert, N. Expression of a functional hyperpolarization-activated current (Ih) in the mouse nucleus reticularis thalami. J. Neurophysiol. 2006, 95, 3073–3085. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Piet, R.; Boehm, U.; Herbison, A.E. Estrous cycle plasticity in the hyperpolarization-activated current ih is mediated by circulating 17β-estradiol in preoptic area kisspeptin neurons. J. Neurosci. 2013, 33, 10828–10839. [Google Scholar] [CrossRef] [Green Version]
  78. Agmon, A.; Connors, B.W. Thalamocortical responses of mouse somatosensory (barrel) cortex in vitro. Neuroscience 1991, 41, 365–379. [Google Scholar] [CrossRef]
  79. Pape, H.C. Queer current and pacemaker: The hyperpolarization-activated cation current in neurons. Annu. Rev. Physiol. 1996, 58, 299–327. [Google Scholar] [CrossRef]
  80. Santoro, B.; Chen, S.; Luthi, A.; Pavlidis, P.; Shumyatsky, G.P.; Tibbs, G.R.; Siegelbaum, S.A. Molecular and functional heterogeneity of hyperpolarization-activated pacemaker channels in the mouse CNS. J. Neurosci. 2000, 20, 5264–5275. [Google Scholar] [CrossRef] [Green Version]
  81. Ventura, C.M.; Kalluri, R. Enhanced Activation of HCN Channels Reduces Excitability and Spike-Timing Regularity in Maturing Vestibular Afferent Neurons. J. Neurosci. 2019, 39, 2860–2876. [Google Scholar] [CrossRef]
  82. Ghasemi, A.; Zahediasl, S. Normality tests for statistical analysis: A guide for non-statisticians. Int. J. Endocrinol. Metab. 2012, 10, 486–489. [Google Scholar] [CrossRef] [PubMed]
Figure 1. (A) DEX did not affect the resting membrane potential (RMP, control: −58.2 [−59.7–−56.9] mV) over the concentration range from 1 µM to 100 µM after 45 min of exposition, as seen in the left panel. In contrast, the HCN channel blocker ZD7288 (40 µM) hyperpolarized every cell exposed (n = 5) and this effect persisted after coapplication with 10 µM DEX, as seen in the middle panel. For analysis of active and passive biophysical membrane properties, currents from −90 pA to +360 pA (increments of 10 pA) were injected into TC relay neurons, with the corresponding current protocol shown above. On the right, representative voltage traces under the baseline conditions are recorded in the current-clamp mode (for a better overview, only every third voltage trace is included). The RMP was derived from the 0 pA current injection step. Δ: absolute change compared to control and ns: not significant. (B) DEX significantly increased neuronal input resistance at a hyperpolarizing current injection of −90 pA, beginning at a concentration of 10 µM and indicating a decrease in cell leakiness during membrane hyperpolarization. Control: 343.4 (252.3–365.6) MΩ, in the left panel. In the middle panel, the effects of ZD7288 on native recordings and after coapplication with 10 µM DEX are shown. ZD7288 induced a strong increase in neuronal input resistance in all cells. Representative voltage traces under baseline conditions and in the presence of 10 µM DEX (blue), as well as 40 µM ZD7288 (orange), are depicted on the right. The input resistance was derived from the voltage changes following a current injection of −90 pA. The accompanying current protocol is shown above and * p < 0.05. (C) Current-voltage relationships were created by injecting currents from −90 pA to +360 pA (10 pA increments) into TC neurons and the change (Δ V) in membrane potential was recorded. The resting membrane potential was derived from Δ V = 0 mV. Significant increases of Δ V were only reliably observable in presence of 100 µM DEX over the complete current range, whereas the effects of 10 µM DEX were restricted to strongly hyperpolarizing current injections. (D) A total of 100 µM DEX shifted the voltage threshold for action potential (AP) generation toward depolarization, while the threshold remained unchanged at lower concentrations. The AP threshold under control conditions was −39.2 (−40.5–−37.6) mV (left panel). In contrast, ZD7288 shifted the AP threshold in all recorded cells toward hyperpolarization, as shown in the middle panel. Representative voltage traces under control conditions and in the presence of 10 µM DEX are depicted on the right. The threshold is indicated by a dotted line. (E) Similarly, only the concentration of 100 µM DEX significantly decreased the frequency (f) of tonic action potential (AP) firing, induced by depolarizing current injections. Corresponding voltage traces under control conditions and after the application of 10 µM DEX are shown on the right.
Figure 1. (A) DEX did not affect the resting membrane potential (RMP, control: −58.2 [−59.7–−56.9] mV) over the concentration range from 1 µM to 100 µM after 45 min of exposition, as seen in the left panel. In contrast, the HCN channel blocker ZD7288 (40 µM) hyperpolarized every cell exposed (n = 5) and this effect persisted after coapplication with 10 µM DEX, as seen in the middle panel. For analysis of active and passive biophysical membrane properties, currents from −90 pA to +360 pA (increments of 10 pA) were injected into TC relay neurons, with the corresponding current protocol shown above. On the right, representative voltage traces under the baseline conditions are recorded in the current-clamp mode (for a better overview, only every third voltage trace is included). The RMP was derived from the 0 pA current injection step. Δ: absolute change compared to control and ns: not significant. (B) DEX significantly increased neuronal input resistance at a hyperpolarizing current injection of −90 pA, beginning at a concentration of 10 µM and indicating a decrease in cell leakiness during membrane hyperpolarization. Control: 343.4 (252.3–365.6) MΩ, in the left panel. In the middle panel, the effects of ZD7288 on native recordings and after coapplication with 10 µM DEX are shown. ZD7288 induced a strong increase in neuronal input resistance in all cells. Representative voltage traces under baseline conditions and in the presence of 10 µM DEX (blue), as well as 40 µM ZD7288 (orange), are depicted on the right. The input resistance was derived from the voltage changes following a current injection of −90 pA. The accompanying current protocol is shown above and * p < 0.05. (C) Current-voltage relationships were created by injecting currents from −90 pA to +360 pA (10 pA increments) into TC neurons and the change (Δ V) in membrane potential was recorded. The resting membrane potential was derived from Δ V = 0 mV. Significant increases of Δ V were only reliably observable in presence of 100 µM DEX over the complete current range, whereas the effects of 10 µM DEX were restricted to strongly hyperpolarizing current injections. (D) A total of 100 µM DEX shifted the voltage threshold for action potential (AP) generation toward depolarization, while the threshold remained unchanged at lower concentrations. The AP threshold under control conditions was −39.2 (−40.5–−37.6) mV (left panel). In contrast, ZD7288 shifted the AP threshold in all recorded cells toward hyperpolarization, as shown in the middle panel. Representative voltage traces under control conditions and in the presence of 10 µM DEX are depicted on the right. The threshold is indicated by a dotted line. (E) Similarly, only the concentration of 100 µM DEX significantly decreased the frequency (f) of tonic action potential (AP) firing, induced by depolarizing current injections. Corresponding voltage traces under control conditions and after the application of 10 µM DEX are shown on the right.
Ijms 24 00519 g001
Figure 2. (A) Relative changes of HCN-mediated Ih current amplitudes at respective membrane potentials after DEX application (10 µM and 100 µM) compared to the control and the effect of the selective HCN channel blocker ZD7288 (40 µM). Beginning at −83 mV, 10 µM DEX induced a significant reduction in Ih. On the right, representative current traces from a TC relay neuron under control conditions and after the application of 10 µM DEX for 45 min. In comparison, 40 µM of ZD7288 resulted in a near complete attenuation of Ih currents over the complete voltage range. The voltage-clamp protocol is shown above and * p < 0.05. (B) Relative reduction in the maximal Ih current amplitude (Ih max) was recorded when the cell was hyperpolarized to −133 mV. The observed changes were significant beginning at a concentration of 10 µM DEX. On the right, representative current traces of Ih max under control conditions and in the presence of 10 µM DEX. The reduction observed after adding 40 µM of ZD7288 (orange) was almost complete, as depicted below. (C) At physiological membrane potentials (−73 mV), the reduction in relative Ih amplitude was only significant after the application of 100 µM DEX. HCN channel blocker ZD7288 reduced Ih−73 mV almost completely in all cells (n = 5). On the right, representative current traces of Ih at −73 mV under control conditions and in the presence of 10 µM DEX. Ns: not significant.
Figure 2. (A) Relative changes of HCN-mediated Ih current amplitudes at respective membrane potentials after DEX application (10 µM and 100 µM) compared to the control and the effect of the selective HCN channel blocker ZD7288 (40 µM). Beginning at −83 mV, 10 µM DEX induced a significant reduction in Ih. On the right, representative current traces from a TC relay neuron under control conditions and after the application of 10 µM DEX for 45 min. In comparison, 40 µM of ZD7288 resulted in a near complete attenuation of Ih currents over the complete voltage range. The voltage-clamp protocol is shown above and * p < 0.05. (B) Relative reduction in the maximal Ih current amplitude (Ih max) was recorded when the cell was hyperpolarized to −133 mV. The observed changes were significant beginning at a concentration of 10 µM DEX. On the right, representative current traces of Ih max under control conditions and in the presence of 10 µM DEX. The reduction observed after adding 40 µM of ZD7288 (orange) was almost complete, as depicted below. (C) At physiological membrane potentials (−73 mV), the reduction in relative Ih amplitude was only significant after the application of 100 µM DEX. HCN channel blocker ZD7288 reduced Ih−73 mV almost completely in all cells (n = 5). On the right, representative current traces of Ih at −73 mV under control conditions and in the presence of 10 µM DEX. Ns: not significant.
Ijms 24 00519 g002
Figure 3. (A) In order to analyze voltage-dependent channel gating and steady-state activation curves of HCN channels under control conditions and after the application of increasing DEX concentrations, normalized tail current amplitudes (Itail) were fitted to a curve using the Boltzmann function. The tail step was set at −103 mV for each voltage step before returning to −43 mV. V1/2: half-maximal activation voltage. (B) Only the high concentration of 100 µM DEX led to a significant left shift (toward hyperpolarization) in V1/2 compared to the control (−86.6 [−85.2–−88.0] mV). Respective current traces under control conditions and after the application of 10 µM DEX are depicted on the right. Tail currents (Itail) were measured subsequent to the hyperpolarizing voltage step at the fixed “tail” step of −103 mV before returning to −43 mV. Δ: absolute change compared to control. Ns: not significant. * p < 0.05. (C) Representative current traces of Ih at the hyperpolarizing voltage step of −133 mV under control conditions and after 45 min of DEX application (10 µM) are depicted above. By fitting a biexponential function to Ih (as indicated by the dotted line), fast (τfast), and slow (τslow) time constants of time-dependent activation were determined. Only the high concentration of 100 µM DEX resulted in a significant deceleration of the fast activation constant compared to the control (231 [214–273] ms), presented as absolute changes. The slow time constant was not affected (control: 1143 [986–1405] ms).
Figure 3. (A) In order to analyze voltage-dependent channel gating and steady-state activation curves of HCN channels under control conditions and after the application of increasing DEX concentrations, normalized tail current amplitudes (Itail) were fitted to a curve using the Boltzmann function. The tail step was set at −103 mV for each voltage step before returning to −43 mV. V1/2: half-maximal activation voltage. (B) Only the high concentration of 100 µM DEX led to a significant left shift (toward hyperpolarization) in V1/2 compared to the control (−86.6 [−85.2–−88.0] mV). Respective current traces under control conditions and after the application of 10 µM DEX are depicted on the right. Tail currents (Itail) were measured subsequent to the hyperpolarizing voltage step at the fixed “tail” step of −103 mV before returning to −43 mV. Δ: absolute change compared to control. Ns: not significant. * p < 0.05. (C) Representative current traces of Ih at the hyperpolarizing voltage step of −133 mV under control conditions and after 45 min of DEX application (10 µM) are depicted above. By fitting a biexponential function to Ih (as indicated by the dotted line), fast (τfast), and slow (τslow) time constants of time-dependent activation were determined. Only the high concentration of 100 µM DEX resulted in a significant deceleration of the fast activation constant compared to the control (231 [214–273] ms), presented as absolute changes. The slow time constant was not affected (control: 1143 [986–1405] ms).
Ijms 24 00519 g003
Figure 4. (A) TC replay neurons display an inwardly directed rectification of the membrane potential upon hyperpolarizing current injection (−350 pA, 500 ms, and the current protocol is shown on top), termed voltage sag. The voltage sag is unmasked in the presence of Ba2+ (green, 150 µM), on the left panel. Low-threshold calcium spike rebound burst-firing (rebound burst) is followed by an HCN channel-mediated afterdepolarization. In comparison, there is no quantifiable voltage sag in the presence of HCN channel blocker ZD7288 (orange, 40 µM), corresponding to a complete suppression. (B) Exemplar current-clamp recordings in the absence and after 45 min of 10 µM DEX (blue) administration with no significant changes. (C) DEX had no significant effect on voltage sag amplitude compared to the control (86.4 [70.7–92.9] mV). Δ: absolute change compared to the control. Ns: not significant. (D) Compared to control measurements (29.0 [27.0–30.8] ms), the delay of rebound burst-firing was only prolongated in the presence of 100 µM DEX. * p < 0.05. (E) The median number of action potentials (AP) during rebound burst spiking (control: 9.5 [7.3–12.5]) was only significantly reduced after the application of 100 µM DEX. (F) Accordingly, only the high concentration of 100 µM DEX significantly abbreviated the median rebound burst duration (DUR) compared to the control (225.5 [126.5–399.0] ms).
Figure 4. (A) TC replay neurons display an inwardly directed rectification of the membrane potential upon hyperpolarizing current injection (−350 pA, 500 ms, and the current protocol is shown on top), termed voltage sag. The voltage sag is unmasked in the presence of Ba2+ (green, 150 µM), on the left panel. Low-threshold calcium spike rebound burst-firing (rebound burst) is followed by an HCN channel-mediated afterdepolarization. In comparison, there is no quantifiable voltage sag in the presence of HCN channel blocker ZD7288 (orange, 40 µM), corresponding to a complete suppression. (B) Exemplar current-clamp recordings in the absence and after 45 min of 10 µM DEX (blue) administration with no significant changes. (C) DEX had no significant effect on voltage sag amplitude compared to the control (86.4 [70.7–92.9] mV). Δ: absolute change compared to the control. Ns: not significant. (D) Compared to control measurements (29.0 [27.0–30.8] ms), the delay of rebound burst-firing was only prolongated in the presence of 100 µM DEX. * p < 0.05. (E) The median number of action potentials (AP) during rebound burst spiking (control: 9.5 [7.3–12.5]) was only significantly reduced after the application of 100 µM DEX. (F) Accordingly, only the high concentration of 100 µM DEX significantly abbreviated the median rebound burst duration (DUR) compared to the control (225.5 [126.5–399.0] ms).
Ijms 24 00519 g004
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Schwerin, S.; Westphal, C.; Klug, C.; Schneider, G.; Kreuzer, M.; Haseneder, R.; Kratzer, S. Sedative Properties of Dexmedetomidine Are Mediated Independently from Native Thalamic Hyperpolarization-Activated Cyclic Nucleotide-Gated Channel Function at Clinically Relevant Concentrations. Int. J. Mol. Sci. 2023, 24, 519. https://doi.org/10.3390/ijms24010519

AMA Style

Schwerin S, Westphal C, Klug C, Schneider G, Kreuzer M, Haseneder R, Kratzer S. Sedative Properties of Dexmedetomidine Are Mediated Independently from Native Thalamic Hyperpolarization-Activated Cyclic Nucleotide-Gated Channel Function at Clinically Relevant Concentrations. International Journal of Molecular Sciences. 2023; 24(1):519. https://doi.org/10.3390/ijms24010519

Chicago/Turabian Style

Schwerin, Stefan, Catharina Westphal, Claudia Klug, Gerhard Schneider, Matthias Kreuzer, Rainer Haseneder, and Stephan Kratzer. 2023. "Sedative Properties of Dexmedetomidine Are Mediated Independently from Native Thalamic Hyperpolarization-Activated Cyclic Nucleotide-Gated Channel Function at Clinically Relevant Concentrations" International Journal of Molecular Sciences 24, no. 1: 519. https://doi.org/10.3390/ijms24010519

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop