Next Article in Journal
A Tribute to Two Master Teachers of Immunology
Previous Article in Journal
Autoimmune Encephalitis—A Multifaceted Pathology
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Modulating Hyperpolarization-Activated Cation Currents through Small Molecule Perturbations: Magnitude and Gating Control

by
Cheng-Shih Chen
1,
Edmund Cheung So
1,† and
Sheng-Nan Wu
2,3,4,*,†
1
Department of Anesthesia, An Nan Hospital, China Medical University, Tainan 70965, Taiwan
2
School of Medicine, National Sun Yat Sen University College of Medicine, Kaohsiung 804, Taiwan
3
Department of Medical Education & Research, An Nan Hospital, China Medical University, Tainan 70965, Taiwan
4
Department of Physiology, National Cheng Kung University Medical College, Tainan 701, Taiwan
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomedicines 2023, 11(8), 2177; https://doi.org/10.3390/biomedicines11082177
Submission received: 9 June 2023 / Revised: 10 July 2023 / Accepted: 27 July 2023 / Published: 2 August 2023
(This article belongs to the Section Cell Biology and Pathology)

Abstract

:
The hyperpolarization-activated cation current (Ih) exhibits a slowly activating time course of the current (Ih) when the cell membrane is hyperpolarized for an extended duration. It is involved in generating electrical activity in various excitable cells. Numerous structurally distinct compounds or herbal drugs have the potential to impact both the magnitude and gating kinetics of this current. Brivaracetam, a chemical analog of levetiracetam known to be a ligand for synaptic vesicle protein 2A, could directly suppress the Ih magnitude. Carisbamate, an anticonvulsant agent, not only inhibited the Ih amplitude but also reduced the strength of voltage-dependent hysteresis (Hys(V)) associated with Ih. Cilobradine, similar to ivabradine, inhibited the amplitude of Ih; however, it also suppressed the amplitude of delayed-rectifier K+ currents. Dexmedetomidine, an agonist of α2-adrenergic receptor, exerted a depressant action on Ih in a concentration-dependent fashion. Suppression of Ih amplitude was observed when GAL-021, a breathing control modulator, was present at a concentration exceeding 30 μM. Lutein, one of the few xanthophyll carotenoids, was able to suppress the Ih amplitude as well as to depress Hys(V)’s strength of Ih. Pirfenidone, a pyridine derivative known to be an anti-fibrotic agent, depressed the Ih magnitude in a concentration- and voltage-dependent fashion. Tramadol, a synthetic centrally active analgesic, was shown to reduce the Ih magnitude, independent of its interaction with opioid receptors. Various herbal drugs, including ent-kaurane-type diterpenoids from Croton tonkinensis, Ganoderma triterpenoids, honokiol, and pterostilbene, demonstrated efficacy in reducing the magnitude of Ih. Conversely, oxaliplatin, a platinum-based chemotherapeutic compound, was observed to effectively increase the Ih amplitude. Collectively, the regulatory effects of these compounds or herbal drugs on cellular function can be partly attributed to their perturbations on Ih.

1. Introduction

The hyperpolarization-activated cation current, also known as the Ih or funny current (If), plays a crucial role in generating repetitive electrical activity in various types of cells, such as heart cells, neurons, and neuroendocrine or endocrine cells [1,2,3,4,5,6,7,8,9,10]. This specific type of ionic current involves a combined flow of Na+ and K+ ions (Figure 1), demonstrating distinctive ion selectivity. These currents exhibit an inwardly rectifying property whereby their amplitude and activation increase in response to more hyperpolarizing potentials [1,2,4,11,12]. The current can be inhibited by CsCl or ivabradine. Given its tonic activity in resting cells, the activation of this current occurs at the resting membrane potential and primarily results in an inward current carried by Na+ ions. Based on this observation, it can be inferred that some cells do not exhibit a truly resting membrane potential and thus deviate from the traditional notion of cellular resting states [8]. The presence or assumption of Ih, as demonstrated in respective models, is crucial because the inward current it generates induces membrane depolarization, playing a pivotal role in initiating action potentials in excitable cells [1,4,8,11]. Recent studies from a computational model of layer V pyramidal cells have described that the slow kinetics of Ih in response to a long hyperpolarizing step can produce long-lasting, activity-dependent modification of membrane excitability in different excitable cell types [13].
Upon long-lasting inverted triangular ramp voltage (Vramp), the forward and backward amplitudes of Ih were noted to be distinct, reflecting the presence of non-equilibrium voltage-dependent hysteresis (Hys(V)) of Ih (Figure 2) [14,15,16,17]. The Hys(V) property of Ih is considered to serve a role in affecting the overall behaviors of excitable cells, including pituitary GH3 cells. In other words, a shift of ion-channel mode in which the voltage sensitivity in gating the charge movement of the current is dependent on the previous state of the HCN channel involved [14,16]. When the membrane potential of an excitable cell undergoes repolarization or hyperpolarization, specifically during the descending limb of the isosceles-triangular Vramp, the current strength of Ih is relatively small according to the Hys(V) loop (Figure 2). However, during the depolarization of the cell membrane, correspondingly, in the ascending limb of this triangular Vramp, the magnitude of Ih significantly increases, leading to a substantial alteration in the membrane potential. As a result, the depolarization of the cell membrane becomes notably pronounced due to the influence of Ih current.
These ionic currents are attributed to channels known as hyperpolarization-activated cyclic nucleotide-gated (HCN) channels, as demonstrated previously [6,11,18,19]. Several mammalian subtypes, namely HCN1, HCN2, HCN3, and HCN4, have been cloned, as reported previously [11,12,18,19]. These subtypes can combine to form either homo- or heterotetramers, each exhibiting distinct biophysical properties. The activation kinetics of HCN1 channels typically activate more slowly compared to HCN2 channels. HCN1 channels have more negative voltage dependence, meaning that they activate at more hyperpolarized potentials compared to HCN2 channels [11,12]. However, it is important to note that these are general trends, and specific kinetic properties can vary depending on the experimental conditions and cellular context. Functional expression of HCN2, HCN3, or a combination of HCN2 and HCN3 channels has been observed in pituitary GH3 cells or other types of endocrine cells [5,6,7,8]. Considering the significance of Ih (i.e., currents mediated by HCNx) in contributing to the excitability and automaticity of excitable cells [12,13,20], any compounds capable of influencing the magnitude, gating behaviors, or voltage-dependent hysteresis (Hys(V)) of Ih can exert a substantial impact on the functional activities of these excitable cells.
In this review paper, our aim was to present information on recently identified compounds or herbal drugs (refer to Table 1) that have a notable influence on the intensity, gating kinetics, and Hys(V) behavior of the current (known as Ih). These compounds have a distinct and significant effect on Ih, many of which differ from their originally developed targets. The IC50 or EC50 values needed for the regulation of Ih magnitude are illustrated in Table 2. Therefore, their impact on Ih should be considered an additional and important mechanism. It is necessary to further clarify the influence of these compounds on cellular function. These effects could offer new insights and enhance our understanding of the potential use of the HCN channel in combating specific types of diseases [11,21].

2. Compounds That Are Known to Inhibit Ih

2.1. Brivaracetam

Brivaracetam (Brivact®, Brivlera®, (2S)-2-[(4R)-2-oxo-4-propylpyrrolidin-1-yl]butanamide), a chemical analog of levetiracetam, is an orally or intravenously bioavailable racetam derivative with anticonvulsant properties [36,37]. Brivaracetam has also been reported to attenuate pain behavior in a murine model of neuropathic pain [36,38]. Recent investigations have shown that, in addition to the inhibition of voltage-gated Na+ currents, brivaracetam at a concentration greater than 10 μM was also found to suppress the amplitude of Ih in pituitary GH3 cells [22]. GH3 cells are not neurons; they are a cell line derived from rat pituitary tumor. Moreover, according to a simulated firing of action potentials generated from the modeled neuron, the firing frequency and amplitude of action potentials were found to be reduced in the presence of brivaracetam [22]. Therefore, it is possible that besides being a high affinity ligand for synaptic vesicle protein 2A (SV2A) [39], brivaracetam can directly perturb the ionic currents, including Ih, hence showing a potential additional impact on the functional activities of different excitable cells.

2.2. Carisbamate

Carisbamate (RWJ-333369, (RS)-2-(2,3-dihydro-1,4-benzodioxin-6-yl)ethyl carbamate), a bioactive orally administered neuromodulator, has been shown to be beneficial for the treatment of different types of convulsive disorders, including drug-resistant focal epilepsy and partial onset seizure [40,41,42]. Previous studies have reported that carisbamate prevents the development and production of epilepsy-like discharges and exerts neuroprotective effects after epilepticus-like injury [40,43]. Of interest, a recent study reported that carisbamate caused a concentration-dependent decrease in Ih amplitude, with an IC50 value of 38 μM [24]. There was also a marked retardation of the activation time course of Ih in response to a 2-s hyperpolarizing command voltage. The presence of carisbamate also suppressed the Hys(V) strength of Ih activation in response to a long-lasting isosceles-triangular Vramp, suggesting that this drug may interact with the voltage-sensing domains of the HCN channel. Apart from its ability to inhibit voltage-gated Na+ current, carisbamate-mediated changes in the magnitude, gating kinetics, and Hys(V) behavior of Ih may also be of pharmacological or therapeutic relevance [24]. Moreover, the Ih has been shown to be functionally present in heart cells [1,3,10]. Therefore, the carisbamate-mediated inhibition of Ih seen in excitable cells may be responsible for its ability to attenuate the increase of heart rate induced by exposure to organophosphate administration, as described previously [43,44]. Further research is also needed to explore the extent to which carisbamate’s inhibition of Ih contributes to its anticonvulsant effects.

2.3. Cilobradine

Cilobradine (DK-AH269, 2-[(3-bromo-5-isoproxy-2-methylphenyl)methylamino]-N-(2,3-dimethylphenyl)acetamide) has been shown to suppress the activity of HCN channels in mouse sinoatrial node cells [45]. Previous observations have revealed its effectiveness in modifying Ih in pancreatic α- or β-cells, thereby influencing hormone secretion [7,46]. In a recent study [26], the effective IC50 required for the cilobradine-induced inhibition of Ih was estimated to be 3.38 μM, a value that tends to be higher than that for its suppression of HCN channels identified in mouse sinoatrial cells. The presence of cilobradine was also noted to alter the impedance amplitude profile of Ih in response to chirp voltage [26]. Furthermore, cilobradine was able to suppress delayed-rectifier K+ currents (IK(DR)) along with an increase in the inactivation time course of the current. As the cilobradine concentration increased from 1 to 3 μM, the midpoint of the steady-state inactivation curve of IK(DR) was shifted along the voltage axis towards hyperpolarizing voltage by approximately 7 mV with no change in the gating charge of the curve during exposure to 1 or 3 μM cilobradine [30]. It is also important to note that exposure to cilobradine has been previously reported to modulate balance function, given that it may concertedly influence functional HCN channels in vestibular hair cells of the inner ear [47] and the KV3.1 channels, which are enriched in the auditory pathway [48]. In other words, the presence of cilobradine may synergistically act on Ih and IK(DR) to influence the functional activities of excitable cells.

2.4. Dexmedetomidine

Dexmedetomidine (Precedex®, (S)-4-[1-(2,3-dimethylphenyl)ethyl]-1H-imidazole), a lipophilic imidazole derivative, is viewed as a potent and selective agonist of the α2-adrenergic receptor [49,50,51]. Previous studies have revealed that this drug exerts a variety of actions on the human brain, such as sedation, anesthetic-sparing effects, and analgesia [52,53,54]. However, there is evidence to highlight the notion that direct interactions with membrane ionic channels may be an unidentified but important mechanism underlying dexmedetomidine-induced action in central neurons [28,53,55,56,57]. In particular, in pituitary GH3 cells, dexmedetomidine produced a depressant action on Ih in a concentration- and time-dependent fashion, with an IC50 or KD value of 1.21 or 1.97 μM, respectively [28]. Cell exposure to dexmedetomidine shifted the steady-state activation curve of Ih toward a more hyperpolarized potential. This drug also diminished the Hys(V) strength of Ih during a long-lasting triangular Vramp. In pheochromocytoma PC12 cells, the presence of dexmedetomidine also suppressed Ih effectively [28].
It is important to highlight the time-dependent effect of dexmedetomidine on Ih. When cells were exposed to dexmedetomidine, it was observed that the time course of Ih activation during prolonged hyperpolarizing pulses slowed down. These findings suggest that the molecule has a higher affinity for the open state of HCN channels, specifically during sustained hyperpolarization, than for closed or resting channels in GH3 cells. This implies that, besides its known ability to bind to α2-adrenergic receptors, dexmedetomidine may directly influence the activation process of the HCN channel, thereby modifying the magnitude and kinetics of Ih in response to prolonged membrane hyperpolarization. Additionally, the blockade of Ih induced by dexmedetomidine could serve as a significant ionic mechanism that effectively reduces the intrinsic membrane excitability of neurons, as well as neuroendocrine or endocrine cells, in vivo [7,28,52,53,54,55,58].
However, it needs to be noted that the sedative properties of dexmedetomidine and its effects on the thalamocortical network might not be primarily influenced by the direct inhibition of Ih. This implies that dexmedetomidine likely acts through other mechanisms, such as α2-adrenoceptor activation and modulation of noradrenergic excitation, to exert its sedative effects [53].

2.5. GAL-021tion

GAL-021 (N2-methoxy-N2-methyl-N4,N6-dipropyl-1,3,5-triazine-2,4,6-triamine or N-[4,6-bis-n-propylamino-(1,3,5)-triazin-2-yl]-N,O-dimethylhydroxyamine) has been developed as a novel breathing control modulator thought to preserve respiratory drive and to protect patients from the respiratory impairment resulting from opioids and other modalities [59]. Previous studies have reported that this agent is an experimental drug demonstrated to inhibit Ca2+-activated K+ channels with large conductance functionally expressed on type 1 cells of the carotid bodies [59,60].
Consistent with previous reports [60], recent findings have shown the ability of GAL-021 to suppress depolarization-evoked Ca2+-activated K+ currents in GH3 cells [30]. However, GAL-021 at a concentration greater than 30 μM was found to inhibit the amplitude of Ih elicited by long-lasting membrane hyperpolarization [30]. If both Ca2+-activated K+ current and Ih are inhibited simultaneously, their individual inhibitory and excitatory effects on neuronal excitability are counteracted. The inhibition of Ca2+-activated K+ currents reduce the hyperpolarizing influence, while the inhibition of Ih reduces the depolarizing influence. It has also been reported that active respiratory neurons express functional HCN channels [61]. Consequently, in addition to the known inhibition of Ca2+-activated K+ currents [30], the actions of GAL-021 on excitable cells may partly result from the suppression of Ih amplitude.

2.6. Lutein

Lutein (3,3′-dihydroxy-α-carotene-6,6′-diene), a xanthophyll carotenoid known as β,ε-carotene-3,3′-diol, is derived from a hydride of a (6R)-β,ε-carotene. It is found in various vegetables and fruits, but notably, it is present in high concentrations in the macula of the human retina, where it acts as a yellow filter [62]. It is a pigment that belongs to the carotenoid family, and its yellow color allows it to selectively absorb certain wavelengths of light. Specifically, lutein absorbs blue and ultraviolet light while allowing other wavelengths, including yellow and longer visible light, to pass through. Recent studies provide evidence that the dietary intake of lutein can lead to the accumulation of lutein in retinal neural tissue, thereby potentially promoting eye and brain health [62,63]. Of interest, a recent report showed that as pituitary GH3 lactotrophs were exposed to lutein, the magnitude of Ih can be inhibited in a concentration-, state-, voltage-, and Hys(V)-dependent manner [32]. The IC50 value required for the inhibition of Ih was 4.1 μM [32]. These results reflect that, besides its antioxidative or anti-inflammatory properties, the presence of lutein can inhibit the magnitude of Ih as well as alter gating and Hys(V) behavior. The lutein’s action would engage in the modifications of spontaneous action potentials present in excitable cells (e.g., GH3 cells), presuming that similar in vivo observations occur.
The functional expression of HCN2, HCN3, or a combination of HCN2 and HCN3 channels, has been reported in GH3 cells [6,8]. Therefore, it seems unlikely that the lutein-induced inhibition of Ih in native cells is specific to a particular isoform. However, there is a possibility that lutein’s blockage of HCN channels may be related to alterations in phosphene perception in the retina [64]. Why lutein in physiological concentrations induces phosphene perception remains to be further studied. However, the observed effect of lutein on specific ionic currents has the potential to contribute to the beneficial effects of lutein in retinal conditions, particularly macular degeneration [65,66]. It would be worthwhile to further investigate whether lutein exhibits selectivity towards different HCN channel isoforms.

2.7. Pirfenidone

Pirfenidone (Esbriet®, 5-methyl-1-phenyl-2(1H)-pyridinone), a pyridine derivative, is thought to act by interfering with the production of transforming growth factor-β and tumor necrosis factor-α [67]. Mounting evidence has shown the effectiveness of pirfenidone either in treating idiopathic pulmonary fibrosis, or in non-small cell lung cancer [67,68]. A previous report showed that the presence of pirfenidone can inhibit the amplitude of Ih in a concentration- and voltage-dependent fashion [34]. Additionally, when GH3 cells were exposed to pirfenidone, the activation time course of Ih became slower in response to sustained membrane hyperpolarization. These findings suggest that the blocking effect of pirfenidone on Ih is not immediate but develops with time after the HCN channel opens, leading to a significant delay in current activation. In addition, exposure to pirfenidone resulted in the suppression of the Hys(V) strength of Ih, which was elicited by a long-lasting triangular Vramp. This suggests that pirfenidone or compounds with similar structures could bind to the open state of the channel and/or inhibit prolonged channel opening [34]. Therefore, this study provides evidence that pirfenidone has the potential to modify specific ionic currents. Such modifications could have implications for therapeutic applications, particularly when pirfenidone is applied to different excitable cells. These findings suggest that pirfenidone might yield additional beneficial effects in various contexts.

2.8. Tramadol

Tramadol ((±)-cis-2-[(dimethylamino)methyl]-1-(3-methoxyphenyl)cyclohexanol) is a synthetic centrally active analgesic, and its clinical use is rapidly increasing. The mechanism of its analgesic actions was thought to feature mixed μ-opioid and non-opioid activity [69]. Of particular interest, increasing evidence has emerged that this drug may be a direct modulator of ion channels that include HCN channels [23]. The presence of tramadol produced a block of Ih in a time- and concentration-dependent manner. This drug at a concentration of 10 μM could shift the activation curve of Ih to more negative potentials, with no change in the slope’s steepness of the curve. Tramadol reduced the firing of spontaneous action potentials in GH3 cells, indicating the tonic activity of Ih in non-voltage-clamped cells [23]. Thus, the direct blockade of Ih by tramadol may partially contribute to the rhythmic activity of neurons or neuroendocrine cells, and similar results are observed in vivo [11,70].

3. Herbal Drugs That Are Known to Inhibit Ih

3.1. Ent-Kaurane-Type Diterpenoids (i.e., Croton-01, Croton-02, Croton-03) from Croton Tonkinensis

The genus croton (Euphorbiaceae) includes about 300 species that are distributed throughout tropical regions. C. tonkinensis Gagnep is a tropic shrub native to northern Vietnam and has been used to exert anti-inflammatory and cancer chemopreventive activities [71]. Earlier reports have shown that the compounds purified from croton could modify different types of ion channels [25,72]. For example, the presence of croton-03 (ent-1β-acetoxy-7α,14β-dihydroxykaur-16-en-15-one) has been shown to suppress the Ih amplitude in pituitary GH3 cells and INS1 insulin-secreting cells [25]. The hysteretic strength of Ih elicited by triangular Vramp was effectively attenuated by adding croton-03. In current-clamp potential recordings, the amplitude of the sag potential in response to long-lasing hyperpolarizing stimuli was also suppressed by the croton-03 presence [25]. Croton-03 also shifted the activation curve of Ih to a more hyperpolarized potential, with no change in the gating charge of the curve. The sag potential is associated with the activation of HCN channels. These channels generate a specific current called the Ih. The sag potential and the underlying Ih current play important roles in regulating membrane excitability and rhythmic activities [25]. Regarding the steady-state activation curve of Ih during exposure to croton-03, the voltage for half-maximal activation was found to be in the range of the firing of action potentials. Furthermore, the presence of croton-03 resulted in a slower activation time course of Ih at different voltage levels [9]. This suggests that the croton-03 molecule has a higher affinity for the open state of HCN channels compared to the closed or resting state of the channels found in GH3 or INS-1 cells. As a result, the degree of Ih blockage caused by croton-03 seems to vary depending on the applied voltage. Moreover, any alterations in Ih amplitude and gating due to croton-03 would be influenced by factors such as the concentration of croton-03, the occurrence of action potentials, and the preexisting resting potential.

3.2. Ganoderma Triterpenoids (Active Constituents of Ganoderma Spores)

Ganoderma mushrooms (Lingzhi in Chinese, or Reishi in Japanese) are a traditional Chinese herbal medicine that has been used as a nutritional supplement [73]. The triterpenoid fraction of Ganoderma, consisting of more than 300 lanostane-tetracyclic compounds, has been shown to be effective at exerting various biological actions, such as that known either to provide antioxidant activities or to produce neuroprotective effects [74,75]. Recent work has demonstrated the ability of Ganoderma triterpenoids to suppress the magnitude and alter the gating kinetics of Ih. The results reflect that these triterpenoids can modify a dose-, time-, and state-dependent activation of Ih in GH3 cells and in HL-1 cardiomyocytes [27]. The IC50 value required for the Ganoderma triterpenoids-mediated block of Ih was estimated to be 11.7 μg/mL. These triterpenoids were reported to contain various nucleosides, including adenosine [27,75]. However, the triterpenoid-mediated inhibition was not reversed by further addition of adenosine receptor antagonists [27]. The current-clamp voltage recordings were also found to decrease the firing of spontaneous action potentials and the magnitude of sag potentials in GH3 cells [27]. It also appears unlikely that the triterpenoid-induced inhibition of Ih in GH3 cells results from nucleosides (e.g., adenosine) possibly contained in their ingredients. The inhibitory action on Ih caused by Ganoderma triterpenoids may thus have a profound impact on the electrical behaviors of excitable cells (e.g., endocrine, and heart cells) if similar in vitro or in vivo findings occur.

3.3. Honokiol

Honokiol (3′,5-di-(2-propenyl)-1,1′-biphenyl-2,2′-diol) is a hydroxylated biphenyl compound obtained from Magnolia officinalis and from other species of the family Magnoliaceae, and has been used in traditional Asian medicines (Hou p’u or Saiboku-tu(o)) [76]. Honokiol is recognized as a potential natural compound that has been shown to exert multiple effects on various cellular responses in different cancer models [77]. Previous investigations have also shown the ability of magnolia bark or honokiol to modify the secretion of catecholamines from the adrenal medulla [78]. At the cellular level, honokiol, or magnolol has been shown to induce Ca2+ mobilization in cortical neurons and neuroblastoma cells [79]. Recent work has also shown that honokiol or other similar structural compounds can interact with the HCN channels to alter the magnitude and gating of Ih during the long-lasting hyperpolarization step in excitable cells, namely GH3 cells and Rolf B1.T olfactory neurons [29]. The steady-state activation curve of Ih in GH3 cells shifted toward a negative voltage in the presence of honokiol. However, the lack of an effect on the gating charge of the curve occurred, reflecting that the honokiol action on the channel might act as a gate to open the channel but not, instead, act on the region that senses the transmembrane potential. Honokiol also suppress Hys(V)’s strength of Ih elicited during triangular Vramp. This compound was also found to suppress the firing of spontaneous action currents measured under cell-attached current recordings in GH3 cells, and this action is thought to be mediated largely by inhibitory action on Ih [29]. Therefore, the inhibition of Ih was noted to be rapid in onset and is therefore likely to be responsible for its modulatory action on the functional activities of sensory neurons or endocrine cells.

3.4. Pterostilbene

Pterostilbene (3′,5′-dimethoxy-4-stilbenol) is a natural demethylated analog of resveratrol and was named after a natural phenolic compound found in Pterocarpus marsupium Roxb (Fabaceae), which is native to India, Nepal, and Sri Lanka. It is one of the active compounds in the extracts of P. marsupium that was used in Ayurvedic medicine for the treatment of various cancers (Ahmad and Rajagopal, 2015). The evidence has shown its inhibitory effects on almost every cellular event that promotes tumor progression toward metastasis in an apoptosis-dependent or apoptosis-independent manner [80,81,82]. Recent evidence has shown that the application of pterostilbene to GH3 cells resulted in the inhibition of Ih in a concentration-dependent manner with an IC50 of 0.84 μM [31]. The presence of pterostilbene increased the activation time constant of Ih elicited by long-lasting membrane hyperpolarization. During exposure to 1 μM pterostilbene, the steady-state activation curve of Ih was distinctly shifted to more hyperpolarizing potentials by about 11 mV, producing channel opening at more negative voltages. However, like resveratrol [83], pterostilbene can stimulate the magnitude of Ca2+-activated K+ currents in pituitary GH3 cells and in hippocampal mHippoE-14 neurons [31]. Therefore, its modifications on ion-channel activity could conceivably be one of the ionic mechanisms underlying pterostilbene-mediated actions, if similar in vitro or in vivo results can emerge in neurons, and in neuroendocrine or endocrine cells.

4. The Compound That Is Known to Stimulate Ih

Oxaliplatin

Oxaliplatin (Eloxatin®, cis-[oxalato(1R,2R-diaminocyclohexane)platinum(II)]) belongs to a family of platinum-based chemotherapeutic compounds. In combination with 5-fluorouracil, this drug has been used in the treatment of advanced colorectal or gastric cancers [84,85]. Despite its good safety profile, its use has been found to confer susceptibility to peripheral neuropathy, affecting sensory and motor nerve fibers, explaining the unsuitability for long-term treatment [21,86]. Recent investigations have revealed the ability of oxaliplatin to modify the magnitude of membrane ionic currents, including Ih [21,34,35]. A previous report showed that the presence of oxaliplatin can exert dual stimulatory actions on two types of ionic currents, namely Ih and membrane electroporation-induced current (IMEP). Unlike those of Ih, the biophysical properties of macroscopic IMEP are virtually stochastic and not yet deterministic. Oxaliplatin-induced stimulation of Ih could be found in pituitary GH3 and R1220 cells and in Roif B1.T olfactory sensory neurons [33]. Rat pituitary R1220 cells were supplied by ScienCell Research Laboratories (Carlsband, CA, USA), https://sciencellonline.com/rat-pituitary-cells, accessed on 1 March 2023.
It is important to note that the oxaliplatin concentration used to block Ih is closely similar to that achieved in the plasma of treated patients (i.e., 3.6–5.6 μM) [87]. The stimulation by oxaliplatin of Ih observed in GH3 cells was not instantaneous and occurred in a time- and concentration-dependent fashion. Moreover, the effects of oxaliplatin on membrane ionic currents were noted to be rapid in onset [33], and they can thus be upstream of the formation of platinum-DNA adducts occurring inside the nucleus [88]. Moreover, the inhibition of Ih caused by exposure to either dexmedetomidine, lutein, pirfenidone, or ent-kaurane-type diterpenoids from C. tonkinensis can be effectively counteracted by the subsequent addition of oxaliplatin [25,26,89,90].

5. Conclusions

As described above, and in published studies, experimental observations have also revealed that a variety of compounds or herbal drugs may directly modulate the magnitude of Ih. Table 3 shows the two-dimensional chemical structures of the compounds presented herein. The modifications of Ih induced by these compounds can impact heart rate, improve neuropathic pain, and demonstrate anti-convulsant effects (Figure 3). Furthermore, owing to the slow activation properties of Ih, which lacks an inactivation process, the time constant for current activation can occasionally extend to around one second, suggesting that this time is closely related to the synaptic delay. The synaptic delay, which refers to the brief period of time it takes for an electrical signal to travel across a synapse, usually ranges from about 0.3 to 5 ms depending on the specific synapse. Several important HCN modulating compounds, including endogenous cytokines, have also been reported, including EC18 [91], and different derivatives of ivabradine like zatebradine/cilobradine [45] or clonidine [92]. EC18 is an important lead structure due to its moderate selectivity. Furthermore, HCN is influenced by endogenous cytokines, such as interferons and interleukins, which could contribute to indirect modulation pathways of the HCN channels [93]. Therefore, when the cells or tissues studied have functional expression of HCNx channels, modifying the magnitude and/or gating properties of Ih will affect the release of neurotransmitters from the presynaptic neuron, thereby influencing synaptic transmission (Figure 3) [11,19,21,70]. Furthermore, the development of compounds that are isoform-specific for HCNx channels would result in modality-specific treatments. Since many herbal drugs do indeed have significant effects on ion channels, attempting to extract and purify these herbal drugs will also be an important issue in future drug development.

Author Contributions

Conceptualization, C.-S.C., S.-N.W. and E.C.S.; methodology, S.-N.W.; software, S.-N.W.; validation, C.-S.C., S.-N.W. and E.C.S.; formal analysis, C.-S.C. and S.-N.W.; investigation, C.-S.C., S.-N.W. and E.C.S.; resources, S.-N.W.; data curation, C.-S.C., S.-N.W. and E.C.S.; writing—original draft preparation, C.-S.C. and S.-N.W.; writing—review and editing, S.-N.W. and E.C.S.; visualization, S.-N.W. and E.C.S.; supervision, S.-N.W.; project administration, C.-S.C. and S.-N.W.; funding acquisition, S.-N.W. and E.C.S. All authors have read and agreed to the published version of the manuscript.

Funding

The research work that led to the composition of this paper was aided in prat by grans by the National Science and Technology Council (NSTC-110-2320-B-006-028, NSTC-111-2314-B-006-028, and NSTC-112-2923-B-006-001) to S.-N.W.). The study was designed, the data were collected, analyzed, and interpreted, the manuscript was written, and the decision to publish the results was made independently. The authors did not play any role in any of these aspects of the study.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original data are available upon reasonable request to the corresponding author.

Acknowledgments

The authors gratefully acknowledge Hsin-Yen Cho and Meng-Cheng Yu for their contribution to the earlier experiments. Due to space constraints, it is not possible to cite all the relevant papers individually, and we apologize for this limitation.

Conflicts of Interest

No conflict of interest, financial or otherwise, are declared by the author(s). The content and writing of this paper are solely the responsibility of the authors.

Abbreviations

HCN channelhyperpolarization-activated cyclic nucleotide-gated channel
Hys(V)voltage-dependent hysteresis
Vrampramp voltage
Ihhyperpolarization-activated cation current

References

  1. Irisawa, H.; Brown, H.F.; Giles, W. Cardiac pacemaking in the sinoatrial node. Physiol. Rev. 1993, 73, 197–227. [Google Scholar] [CrossRef] [PubMed]
  2. Simasko, S.M.; Sankaranarayanan, S. Characterization of a hyperpolarization-activated cation current in rat pituitary cells. Am. J. Physiol. 1997, 272 Pt 1, E405–E414. [Google Scholar] [CrossRef] [PubMed]
  3. Simasko, S.M.; Sankaranarayanan, S. Pacemaker mechanism of porcine sino-atrial node cells. J. Smooth Muscle Res. 2003, 39, 195–204. [Google Scholar]
  4. DiFrancesco, D. Serious workings of the funny current. Prog. Biophys. Mol. Biol. 2006, 90, 13–25. [Google Scholar] [CrossRef]
  5. Kretschmannova, K.; Gonzalez-Iglesias, A.E.; Tomic, M.; Stojilkovic, S.S. Dependence of hyperpolarisation-activated cyclic nucleotide-gated channel activity on basal cyclic adenosine monophosphate production in spontaneously firing GH3 cells. J. Neuroendocrinol. 2006, 18, 484–493. [Google Scholar] [CrossRef]
  6. Kretschmannova, K.; Kucka, M.; Gonzalez-Iglesias, A.E.; Stojilkovic, S.S. The expression and role of hyperpolarization-activated and cyclic nucleotide-gated channels in endocrine anterior pituitary cells. Mol. Endocrinol. 2012, 26, 153–164. [Google Scholar] [CrossRef]
  7. Zhang, Y.; Zhang, N.; Gyulkhandanyan, A.V.; Xu, E.; Gaisano, H.Y.; Wheeler, M.B.; Wang, Q. Presence of functional hyperpolarisation-activated cyclic nucleotide-gated channels in clonal alpha cell lines and rat islet alpha cells. Diabetologia 2008, 51, 2290–2298. [Google Scholar] [CrossRef] [Green Version]
  8. Stojilkovic, S.S.; Tabak, J.; Bertram, R. Ion channels and signaling in the pituitary gland. Endocr. Rev. 2010, 31, 845–915. [Google Scholar] [CrossRef] [Green Version]
  9. Kuo, P.C.; Kao, Z.H.; Lee, S.W.; Wu, S.N. Effects of Sesamin, the Major Furofuran Lignan of Sesame Oil, on the Amplitude and Gating of Voltage-Gated Na(+) and K(+) Currents. Molecules 2020, 25, 3062. [Google Scholar] [CrossRef]
  10. Wu, S.-N.; Fang, Y.-H.; Liu, P.-Y.; Liu, Y.-W. Characterization of Hyperpolarization-Induced Cation Current in Differentiated Human Embryonic Stem Cell-Derived Cardiomyocytes. J. Am. College Cardiol. 2020, 75 (Suppl. 1), 695. [Google Scholar] [CrossRef]
  11. Sartiani, L.; Mannaioni, G.; Masi, A.; Novella Romanelli, M.; Cerbai, E. The Hyperpolarization-Activated Cyclic Nucleotide-Gated Channels: From Biophysics to Pharmacology of a Unique Family of Ion Channels. Pharmacol. Rev. 2017, 69, 354–395. [Google Scholar]
  12. Spinelli, V.; Sartiani, L.; Mugelli, A.; Romanelli, M.N.; Cerbai, E. Hyperpolarization-activated cyclic-nucleotide-gated channels: Pathophysiological, developmental, and pharmacological insights into their function in cellular excitability. Can. J. Physiol. Pharmacol. 2018, 96, 977–984. [Google Scholar] [CrossRef] [Green Version]
  13. Mäki-Marttunen, T.; Mäki-Marttunen, V. Excitatory and inhibitory effects of HCN channel modulation on excitability of layer V pyramidal cells. PLoS Comput. Biol. 2022, 18, e1010506. [Google Scholar] [CrossRef]
  14. Männikkö, R.; Pandey, S.; Larsson, H.P.; Elinder, F. Hysteresis in the voltage dependence of HCN channels: Conversion between two modes affects pacemaker properties. J. Gen. Physiol. 2005, 125, 305–326. [Google Scholar] [PubMed]
  15. Barthel, L.; Reetz, O.; Strauss, U. Use Dependent Attenuation of Rat HCN1-Mediated Ih in Intact HEK293 Cells. Cell. Physiol. Biochem. 2016, 38, 2079–2093. [Google Scholar] [CrossRef]
  16. Fürst, O.; D’avanzo, N. Isoform dependent regulation of human HCN channels by cholesterol. Sci. Rep. 2015, 5, 14270. [Google Scholar] [CrossRef] [Green Version]
  17. Wu, P.M.; Lai, P.C.; ChO, H.Y.; Chuang, T.H.; Wu, S.N.; Tu, Y.F. Effective Perturbations by Phenobarbital on I(Na), I(K(erg)), I(K(M)) and I(K(DR)) during Pulse Train Stimulation in Neuroblastoma Neuro-2a Cells. Biomedicines 2022, 10, 1968. [Google Scholar] [CrossRef]
  18. 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]
  19. Byczkowicz, N.; Eshra, A.; Montanaro, J.; Trevisiol, A.; Hirrlinger, J.; Kole, M.H.; Shigemoto, R.; Hallermann, S. HCN channel-mediated neuromodulation can control action potential velocity and fidelity in central axons. eLife 2019, 8, e42766. [Google Scholar] [CrossRef]
  20. Benedetti, B.; Bieler, L.; Erhardt-Kreutzer, C.; Jakubecova, D.; Benedetti, A.; Reisinger, M.; Dannehl, D.; Thome, C.; Engelhardt, M.; Couillard-Despres, S. Depolarization and Hyperexcitability of Cortical Motor Neurons after Spinal Cord Injury Associates with Reduced HCN Channel Activity. Int. J. Mol. Sci. 2023, 24, 4715. [Google Scholar] [CrossRef]
  21. Dini, L.; Lungo, M.D.; Resta, F.; Melchiorre, M.; Spinelli, V.; Mannelli, L.D.C.; Ghelardini, C.; Laurino, A.; Sartiani, L.; Romanelli, M.N. Selective Blockade of HCN1/HCN2 Channels as a Potential Pharmacological Strategy Against Pain. Front. Pharmacol. 2018, 9, 1252. [Google Scholar] [CrossRef] [Green Version]
  22. Hung, T.-Y.; Wu, S.-N.; Huang, C.-W. The Integrated Effects of Brivaracetam, a Selective Analog of Levetiracetam, on Ionic Currents and Neuronal Excitability. Biomedicines 2021, 9, 369. [Google Scholar] [PubMed]
  23. Liu, Y.-C.; Wang, Y.-J.; Wu, P.-Y.; Wu, S.-N. Tramadol-induced block of hyperpolarization-activated cation current in rat pituitary lactotrophs. Naunyn Schmiedebergs Arch. Pharmacol. 2009, 379, 127–135. [Google Scholar] [CrossRef] [PubMed]
  24. Hung, T.-Y.; Wu, S.-N.; Huang, C.-W. Concerted suppressive effects of carisbamate, an anti-epileptic alkyl-carbamate drug, on voltage-gated Na+ and hyperpolarization-activated cation currents. Front. Cell. Neurosci. 2023, 17, 1159067. [Google Scholar]
  25. Kuo, P.-C.; Liu, Y.-C.; Lo, Y.-C.; Wu, S.-N. Characterization of Inhibitory Effectiveness in Hyperpolarization-Activated Cation Currents by a Group of ent-Kaurane-Type Diterpenoids from Croton tonkinensis. Int. J. Mol. Sci. 2020, 21, 1268. [Google Scholar]
  26. Lu, T.L.; Lu, T.J.; Wu, S.N. Inhibitory Effective Perturbations of Cilobradine (DK-AH269), A Blocker of HCN Channels, on the Amplitude and Gating of Both Hyperpolarization-Activated Cation and Delayed-Rectifier Potassium Currents. Int. J. Mol. Sci. 2020, 21, 2416. [Google Scholar] [CrossRef] [Green Version]
  27. Chang, W.-T.; Gao, Z.-H.; Lo, Y.-C.; Wu, S.-N. Evidence for Effective Inhibitory Actions on Hyperpolarization-Activated Cation Current Caused by Ganoderma Triterpenoids, the Main Active Constituents of Ganoderma Spores. Molecules 2019, 24, 4256. [Google Scholar]
  28. 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 α(2)-Adrenergic Receptors. Int. J. Mol. Sci. 2020, 21, 9110. [Google Scholar] [CrossRef]
  29. Chan, M.H.; Chen, H.H.; Lo, Y.C.; Wu, S.N. Effectiveness in the Block by Honokiol, a Dimerized Allylphenol from Magnolia Officinalis, of Hyperpolarization-Activated Cation Current and Delayed-Rectifier K(+) Current. Int. J. Mol. Sci. 2020, 21, 4260. [Google Scholar] [CrossRef]
  30. Lu, T.L.; Gao, Z.H.; Li, S.W.; Wu, S.N. High Efficacy by GAL-021: A Known Intravenous Peripheral Chemoreceptor Modulator that Suppresses BK(Ca)-Channel Activity and Inhibits I(K(M)) or I(h). Biomolecules 2020, 10, 188. [Google Scholar]
  31. Tan, K.; Chen, P.; Li, S.; Ke, T.; Lin, S.; Yang, C. Characterization of Effectiveness in Concerted I(h) Inhibition and I(K(Ca)) Stimulation by Pterostilbene (Trans-3,5-dimethoxy-4′-hydroxystilbene), a Stilbenoid. Int. J. Mol. Sci. 2020, 21, 357. [Google Scholar]
  32. Chuang, C.W.; Chang, K.P.; Cho, H.Y.; Chuang, T.H.; Yu, M.C.; Wu, C.L.; Wu, S.N. Characterization of Inhibitory Capability on Hyperpolarization-Activated Cation Current Caused by Lutein (β,ε-Carotene-3,3′-Diol), a Dietary Xanthophyll Carotenoid. Int. J. Mol. Sci. 2022, 23, 7186. [Google Scholar] [PubMed]
  33. Chang, W.-T.; Gao, Z.-H.; Li, S.-W.; Liu, P.-Y.; Lo, Y.-C.; Wu, S.-N. Characterization in Dual Activation by Oxaliplatin, a Platinum-Based Chemotherapeutic Agent of Hyperpolarization-Activated Cation and Electroporation-Induced Currents. Int. J. Mol. Sci. 2020, 21, 396. [Google Scholar] [CrossRef] [Green Version]
  34. Chang, W.-T.; Ragazzi, E.; Liu, P.-Y.; Wu, S.-N. Effective block by pirfenidone, an antifibrotic pyridone compound (5-methyl-1-phenylpyridin-2[H-1]-one), on hyperpolarization-activated cation current: An additional but distinctive target. Eur. J. Pharmacol. 2020, 882, 173237. [Google Scholar]
  35. Huang, M.H.; Huang, Y.M.; Wu, S.N. The Inhibition by Oxaliplatin, a Platinum-Based Anti-Neoplastic Agent, of the Activity of Intermediate-Conductance Ca²⁺-Activated K⁺ Channels in Human Glioma Cells. Cell Physiol. Biochem. 2015, 37, 1390–1406. [Google Scholar]
  36. Malawska, B.; Kulig, K. Brivaracetam: A new drug in development for epilepsy and neuropathic pain. Expert Opin. Investig. Drugs 2008, 17, 361–369. [Google Scholar]
  37. Strzelczyk, A.; Klein, K.M.; Willems, L.M.; Rosenow, F.; Bauer, S. Brivaracetam in the treatment of focal and idiopathic generalized epilepsies and of status epilepticus. Expert. Rev. Clin. Pharmacol. 2016, 9, 637–645. [Google Scholar]
  38. Tsymbalyuk, S.; Smith, M.; Gore, C.; Tsymbalyuk, O.; Ivanova, S.; Sansur, C.; Gerzanich, V.; Simard, J.M. Brivaracetam attenuates pain behaviors in a murine model of neuropathic pain. Mol. Pain. 2019, 15, 1744806919886503. [Google Scholar] [CrossRef]
  39. Nicolas, J.; Hannestad, J.; Holden, D.; Kervyn, S.; Nabulsi, N.; Tytgat, D.; Huang, Y.; Chanteux, H.; Staelens, L.; Matagne, A.; et al. Brivaracetam, a selective high-affinity synaptic vesicle protein 2A (SV2A) ligand with preclinical evidence of high brain permeability and fast onset of action. Epilepsia 2016, 57, 201–209. [Google Scholar]
  40. da Silva Fernandes, M.J.; Marques-Carneiro, J.E.; Amorim, R.P.; Araujo MG, L.; Nehlig, A. Neuroprotective agents and modulation of temporal lobe epilepsy. Front Biosci. Elite 2015, 7, 79–93. [Google Scholar]
  41. Strzelczyk, A.; Schubert-Bast, S. Expanding the Treatment Landscape for Lennox-Gastaut Syndrome: Current and Future Strategies. CNS Drugs 2021, 35, 61–83. [Google Scholar] [PubMed]
  42. Elkommos, S.; Mula, M. Current and future pharmacotherapy options for drug-resistant epilepsy. Expert. Opin. Pharmacother. 2022, 23, 2023–2034. [Google Scholar] [CrossRef]
  43. Deshpande, L.S.; Nagarkatti, N.; Ziobro, J.M.; Sombati, S.; DeLorenzo, R.J. Carisbamate prevents the development and expression of spontaneous recurrent epileptiform discharges and is neuroprotective in cultured hippocampal neurons. Epilepsia 2008, 49, 1795–1802. [Google Scholar] [CrossRef] [Green Version]
  44. Deshpande, L.S.; DeLorenzo, R.J. Novel therapeutics for treating organophosphate-induced status epilepticus co-morbidities, based on changes in calcium homeostasis. Neurobiol. Dis. 2020, 133, 104418. [Google Scholar]
  45. Stieber, J.; Wieland, K.; Stöckl, G.; Ludwig, A.; Hofmann, F. Bradycardic and proarrhythmic properties of sinus node inhibitors. Mol. Pharmacol. 2006, 69, 1328–1337. [Google Scholar] [CrossRef] [Green Version]
  46. El-Kholy, W.; Patrick, E.M.; Jocelyn-Manning, F.; Alpana, B.; Xue, T.; Xiaodong, G.; Yi, Z.; Juliane, S.; Ronald, A.L.; Robert, G.T.; et al. Hyperpolarization-activated cyclic nucleotide-gated channels in pancreatic beta-cells. Mol. Endocrinol. 2007, 21, 753–764. [Google Scholar]
  47. Horwitz, G.C.; Risner-Janiczek, J.R.; Jones, S.M.; Holt, J.R. HCN channels expressed in the inner ear are necessary for normal balance function. J. Neurosci. 2011, 31, 16814–16825. [Google Scholar]
  48. Chambers, A.R.; Pilati, N.; Balaram, P.; Large, C.H.; Kaczmarek, L.K.; Polley, D.B. Pharmacological modulation of Kv3.1 mitigates auditory midbrain temporal processing deficits following auditory nerve damage. Sci. Rep. 2017, 7, 17496. [Google Scholar]
  49. Martin, E.; Ramsay, G.; Mantz, J.; Sum-Ping, S.T.J. The role of the alpha2-adrenoceptor agonist dexmedetomidine in postsurgical sedation in the intensive care unit. J. Intensive Care Med. 2003, 18, 29–41. [Google Scholar] [CrossRef]
  50. Lehto, J.; Scheinin, A.; Johansson, J.; Marjamäki, P.; Arponen, E.; Scheinin, H.; Scheinin, M. Detecting a dexmedetomidine-evoked reduction of noradrenaline release in the human brain with the alpha2C-adrenoceptor PET ligand [11C]ORM-13070. Synapse 2016, 70, 57–65. [Google Scholar] [CrossRef]
  51. Lehto, J.; Scheinin, A.; Johansson, J.; Marjamäki, P.; Arponen, E.; Scheinin, H.; Scheinin, M. Dexmedetomidine Exerts a Negative Chronotropic Action on Sinoatrial Node Cells Through the Activation of Imidazoline Receptors. J. Cardiovasc. Pharmacol. 2021, 78, 826–838. [Google Scholar]
  52. Shehabi, Y.; Howe, B.D.; Bellomo, R.; Arabi, Y.M.; Bailey, M.; Bass, F.E.; Bin Kadiman, S.; McArthur, C.J.; Murray, L.; Reade, M.C.; et al. Early Sedation with Dexmedetomidine in Critically Ill Patients. N. Engl. J. Med. 2019, 380, 2506–2517. [Google Scholar] [CrossRef] [PubMed]
  53. 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. 2022, 24, 519. [Google Scholar] [CrossRef]
  54. Chen, Z.; Liu, Z.; Feng, C.; Jin, Y.; Zhao, X. Dexmedetomidine as an Adjuvant in Peripheral Nerve Block. Drug Des. Devel Ther. 2023, 17, 1463–1484. [Google Scholar] [CrossRef]
  55. 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]
  56. Chen, B.-S.; Peng, H.; Wu, S.-N. Dexmedetomidine, an alpha2-adrenergic agonist, inhibits neuronal delayed-rectifier potassium current and sodium current. Br. J. Anaesth. 2009, 103, 244–254. [Google Scholar] [CrossRef] [Green Version]
  57. Kosugi, T.; Mizuta, K.; Fujita, T.; Nakashima, M.; Kumamoto, E. High concentrations of dexmedetomidine inhibit compound action potentials in frog sciatic nerves without alpha(2) adrenoceptor activation. Br. J. Pharmacol. 2010, 160, 1662–1676. [Google Scholar] [CrossRef] [Green Version]
  58. Chiu, K.-M.; Lin, T.-Y.; Lee, M.-Y.; Lu, C.-W.; Wang, M.-J.; Wang, S.-J. Dexmedetomidine protects neurons from kainic acid-induced excitotoxicity by activating BDNF signaling. Neurochem. Int. 2019, 129, 104493. [Google Scholar] [CrossRef]
  59. Golder, F.J.; Dax, S.; Baby, S.M.; Gruber, R.; Hoshi, T.; Ideo, C.; Kennedy, A.; Peng, S.; Puskovic, V.; Ritchie, D.; et al. Identification and Characterization of GAL-021 as a Novel Breathing Control Modulator. Anesthesiology 2015, 123, 1093–1104. [Google Scholar] [CrossRef]
  60. Dallas, M.L.; Peers, C.; Golder, F.J.; Baby, S.; Gruber, R.; MacIntyre, D.E. GAL-021 and GAL-160 are Efficacious in Rat Models of Obstructive and Central Sleep Apnea and Inhibit BKCa in Isolated Rat Carotid Body Glomus Cells. Adv. Exp. Med. Biol. 2015, 860, 361–370. [Google Scholar]
  61. Hawkins, V.E.; Hawryluk, J.M.; Takakura, A.C.; Tzingounis, A.V.; Moreira, T.S.; Mulkey, D.K.; Zhuang, J.; Zang, N.; Ye, C.; Xu, F.; et al. HCN channels contribute to serotonergic modulation of ventral surface chemosensitive neurons and respiratory activity. J. Neurophysiol. 2015, 113, 1195–1205. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  62. Mitra, S.; Rauf, A.; Tareq, A.M.; Jahan, S.; Bin Emran, T.; Shahriar, T.G.; Dhama, K.; Alhumaydhi, F.A.; Aljohani, A.S.; Rebezov, M.; et al. Potential health benefits of carotenoid lutein: An updated review. Food Chem. Toxicol. 2021, 154, 112328. [Google Scholar] [CrossRef] [PubMed]
  63. Johnson, E.J. Role of lutein and zeaxanthin in visual and cognitive function throughout the lifespan. Nutr. Rev. 2014, 72, 605–612. [Google Scholar] [CrossRef]
  64. Bemme, S.; Weick, M.; Gollisch, T. Differential Effects of HCN Channel Block on On and Off Pathways in the Retina as a Potential Cause for Medication-Induced Phosphene Perception. Invest. Ophthalmol. Vis. Sci. 2017, 58, 4754–4767. [Google Scholar] [CrossRef]
  65. Barravecchia, I.; Demontis, G.C. HCN1 channels: A versatile tool for signal processing by primary sensory neurons. Prog. Biophys. Mol. Biol. 2021, 166, 133–146. [Google Scholar] [CrossRef]
  66. Gerhardt, M.J.; Priglinger, S.G.; Biel, M.; Michalakis, S. Biology, Pathobiology and Gene Therapy of CNG Channel-Related Retinopathies. Biomedicines 2023, 11, 269. [Google Scholar] [CrossRef]
  67. Kitano, Y.; Wakimoto, S.; Tamura, S.; Kubota, K.; Domon, Y.; Arakawa, N.; Saito, M.; Sava, B.; Buisson, B. Effects of mirogabalin, a novel ligand for the α₂δ subunit of voltage-gated calcium channels, on N-type calcium channel currents of rat dorsal root ganglion culture neurons. Pharmazie 2019, 74, 147–149. [Google Scholar]
  68. Krämer, M.; Markart, P.; Drakopanagiotakis, F.; Mamazhakypov, A.; Schaefer, L.; Didiasova, M.; Wygrecka, M. Pirfenidone inhibits motility of NSCLC cells by interfering with the urokinase system. Cell Signal 2020, 65, 109432. [Google Scholar] [CrossRef]
  69. Gillen, C.; Haurand, M.; Kobelt, D.J.; Wnendt, S. Affinity, potency and efficacy of tramadol and its metabolites at the cloned human mu-opioid receptor. Naunyn Schmiedebergs Arch. Pharmacol. 2000, 362, 116–121. [Google Scholar] [CrossRef]
  70. Cai, W.; Liu, S.-S.; Li, B.-M.; Zhang, X.-H. Presynaptic HCN channels constrain GABAergic synaptic transmission in pyramidal cells of the medial prefrontal cortex. Biol. Open 2022, 11, bio058840. [Google Scholar] [CrossRef]
  71. Giang, P.M.; Son, P.T.; Lee, J.J.; Otsuka, H. Four ent-kaurane-type diterpenoids from Croton tonkinensis Gagnep. Chem. Pharm. Bull. 2004, 52, 879–882. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  72. Fischer, H.; E Machen, T.; Widdicombe, J.H.; Carlson, T.J.; King, S.R.; Chow, J.W.; Illek, B. A novel extract SB-300 from the stem bark latex of Croton lechleri inhibits CFTR-mediated chloride secretion in human colonic epithelial cells. J. Ethnopharmacol. 2004, 93, 351–357. [Google Scholar] [CrossRef]
  73. Shiao, M.S. Natural products of the medicinal fungus Ganoderma lucidum: Occurrence, biological activities, and pharmacological functions. Chem. Rec. 2003, 3, 172–180. [Google Scholar] [CrossRef] [PubMed]
  74. Chiu, H.-F.; Fu, H.-Y.; Lu, Y.-Y.; Han, Y.-C.; Shen, Y.-C.; Venkatakrishnan, K.; Golovinskaia, O.; Wang, C.-K. Triterpenoids and polysaccharide peptides-enriched Ganoderma lucidum: A randomized, double-blind placebo-controlled crossover study of its antioxidation and hepatoprotective efficacy in healthy volunteers. Pharm. Biol. 2017, 55, 1041–1046. [Google Scholar] [CrossRef] [Green Version]
  75. Ahmad, M.F. Ganoderma lucidum: Persuasive biologically active constituents and their health endorsement. Biomed Pharmacother. 2018, 107, 507–519. [Google Scholar] [CrossRef]
  76. Fujita, M.; Itokawa, H.; Sashida, Y. [Studies on the components of Magnolia obovata Thunb. 3. Occurrence of magnolol and hõnokiol in M. obovata and other allied plants]. Yakugaku Zasshi 1973, 93, 429–434. [Google Scholar] [CrossRef] [Green Version]
  77. Lu, C.-H.; Chen, S.-H.; Chang, Y.-S.; Liu, Y.-W.; Wu, J.-Y.; Lim, Y.-P.; Yu, H.-I.; Lee, Y.-R. Honokiol, a potential therapeutic agent, induces cell cycle arrest and program cell death in vitro and in vivo in human thyroid cancer cells. Pharmacol. Res. 2017, 115, 288–298. [Google Scholar] [CrossRef]
  78. Tachikawa, E.; Takahashi, M.; Kashimoto, T. Effects of extract and ingredients isolated from Magnolia obovata thunberg on catecholamine secretion from bovine adrenal chromaffin cells. Biochem. Pharmacol. 2000, 60, 433–440. [Google Scholar] [CrossRef]
  79. Zhai, H.; Nakade, K.; Mitsumoto, Y.; Fukuyama, Y. Honokiol and magnolol induce Ca2+ mobilization in rat cortical neurons and human neuroblastoma SH-SY5Y cells. Eur. J. Pharmacol. 2003, 474, 199–204. [Google Scholar] [CrossRef]
  80. Benlloch, M.; Obrador, E.; Valles, S.L.; Rodriguez, M.L.; Sirerol, J.A.; Alcácer, J.; Pellicer, J.A.; Salvador, R.; Cerdá, C.; Sáez, G.T.; et al. Pterostilbene Decreases the Antioxidant Defenses of Aggressive Cancer Cells In Vivo: A Physiological Glucocorticoids- and Nrf2-Dependent Mechanism. Antioxid. Redox Signal. 2016, 24, 974–990. [Google Scholar] [CrossRef] [Green Version]
  81. Huang, Y.; Du, J.; Mi, Y.; Li, T.; Gong, Y.; Ouyang, H.; Hou, Y. Long Non-coding RNAs Contribute to the Inhibition of Proliferation and EMT by Pterostilbene in Human Breast Cancer. Front. Oncol. 2018, 8, 629. [Google Scholar] [CrossRef] [PubMed]
  82. Tan, K.; Chen, P.; Li, S.; Ke, T.; Lin, S.; Yang, C. Pterostilbene inhibits lung squamous cell carcinoma growth in vitro and in vivo by inducing S phase arrest and apoptosis. Oncol. Lett. 2019, 18, 1631–1640. [Google Scholar] [CrossRef] [Green Version]
  83. Li, H.-F.; Chen, S.-A.; Wu, S.-N. Evidence for the stimulatory effect of resveratrol on Ca(2+)-activated K+ current in vascular endothelial cells. Cardiovasc. Res. 2000, 45, 1035–1045. [Google Scholar] [CrossRef] [Green Version]
  84. Graham, J.; Mushin, M.; Kirkpatrick, P. Oxaliplatin. Nat. Rev. Drug Discov. 2004, 3, 11–12. [Google Scholar] [CrossRef]
  85. Chen, C.-C.; Chen, L.-T.; Tsou, T.-C.; Pan, W.-Y.; Kuo, C.-C.; Liu, J.-F.; Yeh, S.-C.; Tsai, F.-Y.; Hsieh, H.-P.; Chang, J.-Y. Combined modalities of resistance in an oxaliplatin-resistant human gastric cancer cell line with enhanced sensitivity to 5-fluorouracil. Br. J. Cancer 2007, 97, 334–344. [Google Scholar] [CrossRef] [Green Version]
  86. Kanat, O.; Ertas, H.; Caner, B. Platinum-induced neurotoxicity: A review of possible mechanisms. World J. Clin. Oncol. 2017, 8, 329–335. [Google Scholar] [CrossRef]
  87. Jerremalm, E.; Wallin, I.; Ehrsson, H. New insights into the biotransformation and pharmacokinetics of oxaliplatin. J. Pharm. Sci. 2009, 98, 3879–3885. [Google Scholar] [CrossRef]
  88. Ta, L.E.; Espeset, L.; Podratz, J.; Windebank, A.J. Neurotoxicity of oxaliplatin and cisplatin for dorsal root ganglion neurons correlates with platinum-DNA binding. Neurotoxicology 2006, 27, 992–1002. [Google Scholar] [CrossRef]
  89. Chuang, T.H.; Cho, H.Y.; Wu, S.N. The Evidence for Sparsentan-Mediated Inhibition of I(Na) and I(K(erg)): Possibly Unlinked to Its Antagonism of Angiotensin II or Endothelin Type a Receptor. Biomedicines 2021, 10, 86. [Google Scholar] [CrossRef]
  90. Wu, S.-N.; Wu, C.-L.; Cho, H.-Y.; Chiang, C.-W. Effective Perturbations by Small-Molecule Modulators on Voltage-Dependent Hysteresis of Transmembrane Ionic Currents. Int. J. Mol. Sci. 2022, 23, 9453. [Google Scholar] [CrossRef]
  91. Del Lungo, M.; Melchiorre, M.; Guandalini, L.; Sartiani, L.; Mugelli, A.; Koncz, I.; Szel, T.; Varro, A.; Romanelli, M.N.; Cerbai, E. Novel blockers of hyperpolarization-activated current with isoform selectivity in recombinant cells and native tissue. Br. J. Pharmacol. 2012, 166, 602–616. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  92. 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]
  93. Oniani, T.; Vinnenberg, L.; Chaudhary, R.; Schreiber, J.A.; Riske, K.; Williams, B.; Pape, H.-C.; White, J.A.; Junker, A.; Seebohm, G.; et al. Effects of axonal demyelination, inflammatory cytokines and divalent cation chelators on thalamic HCN channels and oscillatory bursting. Int. J. Mol. Sci. 2022, 23, 6285. [Google Scholar] [CrossRef]
Figure 1. Simplified representation graph of a hyperpolarization-activated cyclic nucleotide-gated (HCN) channel. The image on the right displays a voltage-clamp protocol. The top graph illustrates the voltage command applied for hyperpolarization, while the bottom graph represents the waveform of the hyperpolarization-activated cation current (Ih). This demonstrates the characteristic slow activation of Ih without an accompanying inactivation process during prolonged membrane hyperpolarization. The image on the left depicts the conductance of the HCN channel in its open state. This ion channel is known to exhibit permeability comparable to that of both Na+ and K+ ions. In this state, Na+ ions flow from the extracellular space into the intracellular space, while K+ ions move in the opposite direction, both driven by the electrochemical gradient.
Figure 1. Simplified representation graph of a hyperpolarization-activated cyclic nucleotide-gated (HCN) channel. The image on the right displays a voltage-clamp protocol. The top graph illustrates the voltage command applied for hyperpolarization, while the bottom graph represents the waveform of the hyperpolarization-activated cation current (Ih). This demonstrates the characteristic slow activation of Ih without an accompanying inactivation process during prolonged membrane hyperpolarization. The image on the left depicts the conductance of the HCN channel in its open state. This ion channel is known to exhibit permeability comparable to that of both Na+ and K+ ions. In this state, Na+ ions flow from the extracellular space into the intracellular space, while K+ ions move in the opposite direction, both driven by the electrochemical gradient.
Biomedicines 11 02177 g001
Figure 2. Simplified representation graph of the voltage-dependent hysteresis (Hys(V)) of Ih. On the left, a representation diagram of a long-lasting inverted triangle represents a ramp voltage (Vramp, indicated in blue). The dashed arrow indicates the change in voltage over time. On the right is a graph representing the relationship of voltage versus whole-cell Ih current, specifically illustrating the voltage-dependent hysteresis (Hys(V)) (depicted in orange color). The bold dotted lines are positioned at membrane potentials of −40 and −160 mV, respectively. The light dotted curve arrows adjacent to the orange line indicate the counterclockwise direction of the current flow over time. The gray shaded area in the diagram represents the hysteresis strength and is enclosed by the Ih current during both the descending (forward) and ascending (backward) limbs of the triangular Vramp. The Ih current induced by the descending limb of the triangular Vramp is noticeably smaller than the current induced by the ascending limb of the Vramp.
Figure 2. Simplified representation graph of the voltage-dependent hysteresis (Hys(V)) of Ih. On the left, a representation diagram of a long-lasting inverted triangle represents a ramp voltage (Vramp, indicated in blue). The dashed arrow indicates the change in voltage over time. On the right is a graph representing the relationship of voltage versus whole-cell Ih current, specifically illustrating the voltage-dependent hysteresis (Hys(V)) (depicted in orange color). The bold dotted lines are positioned at membrane potentials of −40 and −160 mV, respectively. The light dotted curve arrows adjacent to the orange line indicate the counterclockwise direction of the current flow over time. The gray shaded area in the diagram represents the hysteresis strength and is enclosed by the Ih current during both the descending (forward) and ascending (backward) limbs of the triangular Vramp. The Ih current induced by the descending limb of the triangular Vramp is noticeably smaller than the current induced by the ascending limb of the Vramp.
Biomedicines 11 02177 g002
Figure 3. Changes that occur when the magnitude and gating properties of Ih are modified. Each solid arrow represents a specific type of channel that may occur.
Figure 3. Changes that occur when the magnitude and gating properties of Ih are modified. Each solid arrow represents a specific type of channel that may occur.
Biomedicines 11 02177 g003
Table 1. Compounds presented in this paper that can inhibit the hyperpolarization-activated cation current (Ih), as well as compounds that can stimulate this current.
Table 1. Compounds presented in this paper that can inhibit the hyperpolarization-activated cation current (Ih), as well as compounds that can stimulate this current.
  • Compounds known to inhibit Ih
  a.
Brivaracetam
  b.
carisbamate
  c.
cilobradine
  d.
dexmedetomidine
  e.
GAL-021
  f.
lutein
  g.
pirfenidone
  h.
tramadol
2.
Herbal drugs known to inhibit Ih
  a.
ent-krauane-type diterpenoids
  b.
Ganoderma triterpenoids
  c.
honokiol
  d.
pterostilbene
3.
Compound known to stimulate Ih
  a.
Oxaliplatin
Table 2. Summary showing IC50 or EC50 values for modulating Ih amplitude as indicated in the paper.
Table 2. Summary showing IC50 or EC50 values for modulating Ih amplitude as indicated in the paper.
CompoundIC50Refs.CompoundIC50Refs.
Brivaracetamgreater than 25 μM[22]Tramadol13.6 μM[23]
Carisbamate38 μM[24]Croton-01 *, croton-02 *, and croton-03 *2.89, 6.25, and 2.84 μM[25]
Cilobradine3.38 μM[26]Ganoderma triterpenoids11.7 μg/ml[27]
Dexmedetomidine1.21 μM[28]Honokiol2.1 μM[29]
GAL-021greater than 30 μM[30]Pterostilbene0.84 μM[31]
Lutein4.1 μM[32]Oxaliplatin1.3 μM **[33]
Pirfenidone3.65 μM[34]
* belongs to ent-kaurane-type diterpenoids. ** The value represents the EC50, since oxaliplatin was found to activate the Ih amplitude [35].
Table 3. Two-dimensional chemical structures of the compounds presented in this paper.
Table 3. Two-dimensional chemical structures of the compounds presented in this paper.
CompoundChemical Structure
BrivaracetamBiomedicines 11 02177 i001
CarisbamateBiomedicines 11 02177 i002
CilobradineBiomedicines 11 02177 i003
DexmedetomidineBiomedicines 11 02177 i004
GAL-021Biomedicines 11 02177 i005
LuteinBiomedicines 11 02177 i006
PirfenidoneBiomedicines 11 02177 i007
TramadolBiomedicines 11 02177 i008
Croton-01, croton-02, and croton-03Biomedicines 11 02177 i009
HonokiolBiomedicines 11 02177 i010
PterostilbeneBiomedicines 11 02177 i011
OxaliplatinBiomedicines 11 02177 i012
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

Chen, C.-S.; So, E.C.; Wu, S.-N. Modulating Hyperpolarization-Activated Cation Currents through Small Molecule Perturbations: Magnitude and Gating Control. Biomedicines 2023, 11, 2177. https://doi.org/10.3390/biomedicines11082177

AMA Style

Chen C-S, So EC, Wu S-N. Modulating Hyperpolarization-Activated Cation Currents through Small Molecule Perturbations: Magnitude and Gating Control. Biomedicines. 2023; 11(8):2177. https://doi.org/10.3390/biomedicines11082177

Chicago/Turabian Style

Chen, Cheng-Shih, Edmund Cheung So, and Sheng-Nan Wu. 2023. "Modulating Hyperpolarization-Activated Cation Currents through Small Molecule Perturbations: Magnitude and Gating Control" Biomedicines 11, no. 8: 2177. https://doi.org/10.3390/biomedicines11082177

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