Skip to main content
Current Cardiology Reviews logoLink to Current Cardiology Reviews
. 2022 Jul 14;18(4):e040222200836. doi: 10.2174/1573403X18666220204142436

Review: HCN Channels in the Heart

Anne-Sophie Depuydt 1, Steve Peigneur 1, Jan Tytgat 1,*
PMCID: PMC9893134  PMID: 35125083

Abstract

Pacemaker cells are the basis of rhythm in the heart. Cardiovascular diseases, and in particular, arrhythmias are a leading cause of hospital admissions and have been implicated as a cause of sudden death. The prevalence of people with arrhythmias will increase in the next years due to an increase in the ageing population and risk factors. The current therapies are limited, have a lot of side effects, and thus, are not ideal. Pacemaker channels, also called hyperpolarization-activated cyclic nucleotide-gated (HCN) channels, are the molecular correlate of the hyperpolarization-activated current, called Ih (from hyperpolarization) or If (from funny), that contribute crucially to the pacemaker activity in cardiac nodal cells and impulse generation and transmission in neurons. HCN channels have emerged as interesting targets for the development of drugs, in particular, to lower the heart rate. Nonetheless, their pharmacology is still rather poorly explored in comparison to many other voltage-gated ion channels or ligand-gated ion channels. Ivabradine is the first and currently the only clinically approved compound that specifically targets HCN channels. The therapeutic indication of ivabradine is the symptomatic treatment of chronic stable angina pectoris in patients with coronary artery disease with a normal sinus rhythm. Several other pharmacological agents have been shown to exert an effect on heart rate, although this effect is not always desired. This review is focused on the pacemaking process taking place in the heart and summarizes the current knowledge on HCN channels.

Keywords: HCN channels, hyperpolarization-activated current, automaticity, sinus node dysfunction, atrial fibrillation, ivabradine

1. INTRODUCTION

Cardiovascular diseases result in the death of 17.9 million people every year. Based on data from the World Health Organization (WHO), the total annual number of deaths from noncommunicable diseases (NCDs) is projected to increase to 52 million by 2030 [1]. Due to an increase in the ageing population and risk factors, such as tobacco use, unhealthy diet, physical inactivity, and harmful use of alcohol, among others, it is more important than ever to focus on cardiovascular drugs.

Pacemaker channels or hyperpolarization-activated cyclic nucleotide-gated (HCN) channels, designated as HCN1–4, are the molecular correlate of the hyperpolarization-activated current, called Ih (from hyperpolarization) or If (from funny), that contribute crucially to the pacemaker activity in cardiac nodal cells and impulse generation and transmission in neurons [2, 3]. The name “funny current” arose due to HCN channels displaying a set of specific biophysical properties that set them apart from most of the other ion channels. For instance, HCN channels activate upon hyperpolarization and deactivate upon depolarization. Moreover, their activation curve shifts to more positive potentials following the binding of cAMP to the channel [4]. HCN channels are permeable for both sodium and potassium ions with a reversal potential near -35 mV [5].

Interestingly, the pharmacology of HCN channels is still rather poorly explored compared to many other voltage-gated ion channels or ligand-gated ion channels. Nonetheless, HCN channels have emerged as interesting targets for the development of drugs, in particular, to modulate the heart rate. The properties of atrial and ventricular If, and its modulation by pharmacological interventions, have been the subject of intense study, including the synthesis and characterization of compounds able to preferentially block HCN1, HCN2, or HCN4 isoforms. Altogether, these studies have provided important insights for future pharmacological strategies exploiting the unique properties of this channel family: the prevalence of different HCN subtypes in organs and tissues, the possibility to target HCN gain- or loss-of-function associated with disease, the feasibility of novel isoform-selective drugs, as well as the discovery of HCN-mediated effects for old medicines.

Ivabradine is the first and currently the only clinically approved compound that specifically targets HCN channels [6]. The therapeutic indication of ivabradine is the symptomatic treatment of chronic stable angina pectoris in patients with coronary artery disease and a normal sinus rhythm [7, 8].

This review focuses on the pacemaking process taking place in the heart. We first briefly summarize the current knowledge of the structure and the biophysical properties of HCN channels. We then discuss the molecular mechanisms underlying the regulation, modulation and function of HCN channels, and their role in human diseases by focusing on heart failure (HF) and arrhythmias. Finally, current pharmacology in chronic heart failure will be discussed.

2. STRUCTURE

Four different genes encode for different isoforms (HCN1–4). Topologically, HCN channels are shown to have an overall structure similar to voltage-gated potassium (Kv) channels. Just as in Kv channels [9-11], four HCN subunits assemble to form a functional channel. The subunits can form different homotetramers or heterotetramers with distinct biophysical properties [12, 13]. Each subunit consists of a cytosolic amino (N)-terminus, a core region, and a cytosolic carboxy (C)-terminus. The most important domain is the channel core, composed of six α-helical transmembrane (TM) domains (S1–S6), with two forming the pore (S5–S6) and four forming the voltage sensor domain (S1–S4) [12].

The pore domain (PD) is formed between segments S5 and S6, and contributes to the central pore. The last transmembrane domain is followed by a cytoplasmic region, which includes a cyclic nucleotide–binding domain (CNBD), where cAMP and cGMP can bind to the channel [14]. This domain is highly conserved, in contrast to the C-terminal part of the C-terminal domain, which shows strong sequence variation among the HCN isoforms. The CNBD is connected to the S6 segment via an intracellular C-linker. The binding of cyclic nucleotides to the CNBD induces conformational changes that facilitate channel opening by removing channel inhibition, leading to faster opening kinetics and a shift of the voltage dependence towards more positive values. However, the open probability of HCN can be increased by cAMP, but unlike cyclic nucleotide-gated (CNG) channels, cyclic nucleotides are not a prerequisite for channel opening [5, 12, 15-17]. Moreover, Lee and MacKinnon described another cytoplasmic region, which is unique for HCN. This HCN domain contains 45 amino acids, forming a 3-α-helical domain. It precedes the S1-helix and is wedged between the voltage sensor and the cytoplasmic domains. Since it connects to the S4 helix (near the short S4–S5 linker) from the same subunit and the C-linker and CNBD from an adjacent subunit, it is thought to play a role in the unusual gating behaviour of HCN channels [12].

The S1 to S4 segments form the voltage-sensing domain (VSD), located peripherally to the pore region. The VSDs in HCN1 are “non-swapped”, which means that each voltage-sensing domain contacts the pore through amino acids from the same polypeptide chain (i.e., the same subunit) [12]. This is in contrast to Kv channels (Kv1–Kv9) [18] and voltage-dependent sodium (Nav) and calcium (Cav) channels, where the voltage sensors interact with an adjacent subunit (i.e., they are domain-swapped) [12]. Structural studies using Kv channels have shown that in this canonical model of VSD–PD coupling, the activation energy in the S4 segment of the VSD passes to the S4–S5 linker via the covalent linkage, and subsequently, the S4–S5 linker is formed with the PD via a noncovalent interaction between the S4–S5 linker and the S6 to open the pore gate. Since these channels have domain-swapped VSDs, they must have a long α-helical S4–S5 linker [19, 20]. In HCN1, the S4–S5 linker is much shorter and not α-helical [12].

In addition, the VSD comprises an S4-helix with a regular sequence of positively charged amino acids (lysine and arginine), very similar to depolarization-activated channels [21, 22]. The S4 is thought to function as the voltage sensor by moving charge through the membrane electric field in response to depolarization [23]. In comparison to Kv channels, mammalian HCN channels have a much larger S4 domain, containing 9 basic residues and 1 serine residue at every third position [21]. Regardless of having opposite gating mechanisms, both types have a S4 segment moving upward during depolarization and downward during hyperpolarization [23].

Several studies have proposed that the S4–S5 linker plays an essential role in the HCN channel gating mechanism [12, 24, 25]. However, the exact gating mechanism remains debatable. More recent work suggests that the reversed polarity of HCN channels can be caused by the extraordinary length of the S4 helix. Compared to other voltage-dependent ion channels, the S4 helix of HCN1 contains two additional helical turns on the intracellular side. Because of its length, the S4 segment connects to a C-linker in the cytoplasm to stabilize a closed pore when the voltage sensor is depolarized. In addition, the S4, S5 and S6 helices seem to be closely packed together, enabling the depolarized voltage sensor to stabilize a closed pore. Finally, the HCN domain stabilizes the closed pore in the setting of a depolarized voltage sensor. According to this model, the upward movement of the S4-helix has an inhibitory effect on the pore opening. When the membrane potential is hyperpolarized, the voltage sensor moves downward and disrupts these stabilizing interactions, allowing the S6 helices to spontaneously open [12]. Flynn et al. mutated the S4–S5 linker, the covalent linkage between the VSD and PD [26], and Cowgill et al. systematically developed an array of mosaic channels to display the full voltage-activation phenotypes observed in voltage-gated ion channels [24]. The results showed that hyperpolarization-dependent activation does not require the S4–S5 linker nor the extra length of the HCN S4-helix [24, 26]. In contrast, they did confirm that the gating of HCN channels is inconsistent with the canonical model of VSD–PD coupling in Kv channels [12, 24, 26]. The most recent study by Ramentol et al. made an attempt to fully understand the HCN gating mechanism [25]. They identified two residues that appeared to play a crucial role in the polarity of HCN1, namely W355 (part of the QWE motif in S4) and N370 (S5). When W355 and N370 mutate individually, current conditions are detected and the membrane is depolarized. For both the W355N and N370W channel mutants, a 25% conductance remains at positive voltages. This finding suggests that these mutations destabilize the closed pore, and thus, cause partial opening. Finally, mutation of both the residues (W355N–N370W) could reverse the voltage dependence of HCN channels with no change in the pore structure. This double mutation removes stable interactions, indicating that small differences in the energy of the closed and open states can determine the channel’s polarity [25].

Different regulatory proteins form macromolecular complexes with HCN channel subunits and define HCN channel function. For example, the MinK-related peptide 1 protein (MiRP1; encoded by KCNE2) enhances protein and current expression in the cardiac tissue. MiRP1 can cause an upregulation of the channels and alters the kinetics of the various HCN isoforms [27-29]. Mutations in this KCNE2 gene are related to cardiac disorders, including cardiac arrhythmias [30]. Another accessory subunit interacting with HCN channels is Caveolin-3 (CAV3), which is known to affect the correct function of the HCN4 current in the heart [31, 32] (Fig. 1).

Fig. (1).

Fig. (1)

The location, biophysical properties, transmembrane topology of HCN channels, and regulation of HCN channels by interacting proteins. Figure adapted from Servier Medical Art by Servier. https://smart.servier.com/ (Creative Commons Attribution 3.0 Unported License.)

3. BIOPHYSICAL PROPERTIES

HCN channels are members of the voltage-gated pore-loop channel superfamily and are related to the CNG channels and the Kv channels Kv10–Kv12 [5]. Unique among voltage-gated channels, HCN channels are activated when the membrane potential becomes more negative. Upon activation, the inwardly directed cation current slowly depolarizes pacemaker cells to the threshold for action potential (AP) firing. The mixed Na+ and K+ permeability of HCN channels is essential for their function as membrane-depolarizing channels, yielding a reversal potential of -20 to -30 mV under physiological conditions [33, 34]. Surprisingly, the selectivity filters of HCN channels are similar to Kv channels since the essential amino acids required for selecting K+, the common GYG-motif, are conserved among both the channels [12]. Despite the similarity at the molecular level, the HCN channel has a Na+/K+ permeability ratio of 1:3 to 1:5 [15-17, 34], whereas Kv has a ratio of 1:1000 [35].

In GYG-containing Kv channels, the selectivity filter sequence is ‘T/S-V/I/L/T-GYG’, forming four binding sites that allow potassium ions to pass through the ion pore [36, 37]. Since earlier studies showed that (i) large cations, such as Ba2+ or tetraethylammonium, do not block If, in contrast with Kv channels [17, 38] and (ii) the typical Thr/Ser residue in Kv channels is mutable to Cys (specific to HCN) without affecting the selectivity [36], HCN channels were thought to have a functionally wider pore. A better understanding of the molecular basis of these intriguing biophysical properties was recently obtained by determining the cryogenic electron microscopy (EM) structures of the human HCN1 channel [12]. In this study, the selectivity filter of HCN1 channels was compared to the filter of a distantly related K+ channel, namely the KcsA potassium channel. It was found that in two of the four filters sites, the Tyr-sidechain in the GYG sequence of HCN1 reorientates to 180 degrees. As a result, the carbonyl oxygen atoms of the peptide backbone that would form two ion binding sites to coordinate K+ ions are no longer facing the ion pathway in HCN1, leading to a selectivity filter only preserving two binding sites. Since two tyrosine residues bind to one potassium, the HCN pore will only bind one potassium. Single ion binding pores allow more space, which is kinetically favorable for passing sodium [12].

4. REGULATION AND MODULATION

HCN channels are activated by hyperpolarization of the membrane potential; however, this activation can be influenced by cyclic nucleotides interacting with the CNBD, which is located in the intracellular C-terminal region. cAMP accelerates channel opening and shifts the activation curve to a more positive potential [15-17, 39]. The shift in V1/2 depends on the channel subtype and the conditions. It varies between 10 and 25 mV for HCN2 and HCN4, but only 2–6 mV for HCN1 [40, 41]. The EC50 of cAMP ranges between 0.06 mM and 1.53 mM. Surprisingly, HCN3 is not activated by cAMP; the activation curve is even slightly shifted to more negative voltages [42, 43]. While the HCN channel is inhibited at depolarized potentials, the binding of cAMP removes this inhibition. This induces conformational changes and increases the open probability of the channel pore [40, 44]. Lee and MacKinnon investigated the effect of cAMP by determining the cryo-EM structure of HCN1 in the presence of cAMP. They observed that cAMP binding induces local conformational changes that are propagated from the CNBD via the C-linker to the pore, initiating a rotation of the gate-forming inner helices toward the opening. This could explain why cAMP favours HCN channel opening [12]. The reason for the lack of sensitivity of HCN3 to cAMP, despite the presence of a functional CNBD in the intracellular region, is still unknown. Stieber et al. suggest the rationale of a shorter C-terminal sequence after the CNBD, which alters the normal autoinhibition of the channel [43].

Another full agonist of HCN channels is guanosine-3’,5’-cyclic monophosphate (cGMP). Although cGMP displays a 10-fold lower potency for HCN channels, the shift in V1/2 is in the same range as that of cAMP at saturating concentrations [17]. cGMP and cAMP bind to the CNBD in a similar way, with a different orientation of the purine ring [14]. Another modulator of HCN channels is cytidine-3’,5’-cyclic monophosphate (cCMP), acting as a partial agonist [39]. On HCN2 and HCN4 channels, cCMP shifts the activation curve to more depolarized potentials and enhances current activation and decreases current deactivation. cCMP does not affect HCN1 and HCN3. The voltage shift and the maximal increase in the current amplitude are significantly smaller than those observed for cAMP, corroborating the behaviour of a partial agonist [45]. Uridine-3’,5’-cyclic monophosphate (cUMP), purine-3’,5’-cyclic monophosphate (cPMP), and 2-amino-cPMP shift the activation curve to more positive potentials and increase channel conductance for HCN2, with an efficacy similar to cAMP and cGMP. In contrast, inosine-3’,5’-cyclic monophosphate (cIMP), as well as cCMP, behave as weak activators [46]. In addition to cyclic mononucleotides, cyclic dinucleotides can also modulate HCN channels. In mouse sinoatrial node (SAN) myocytes, they behave as antagonists, being able to reduce Ih [47]. Cyclic [guanosine-(2’-5’)-monophosphate-adenosine-(3’-5’)-monophosph-ate], the cyclic dinucleotide that is found in mammals, caused a shift in the activation curve toward more negative potentials and a 30% reduction of firing rate when 100 µM was added to single SAN cells. For the HCN4 channel expressed in human embryonic kidney (HEK) 293 cells, the dose-response curve yielded an IC50 value of 114 nM. Other cyclic dinucleotides could completely reverse the effect of cAMP on the activation curve; for instance, cyclic di-(3’,5’)-GMP, which is ∼16 times less potent than cyclic [guanosine-(2’-5’)-monophosphate-adenosine-(3’-5’)-monophosphate] (IC50 1.8 mM) [47].

The modulation of HCN channels has been extensively reviewed [48-51]. Besides the cytosolic concentration of cAMP, it is well known that other mechanisms affect channel activity. HCN channel activation also shifts towards depolarizing potentials by phosphoinositides, such as phosphatidylinositol-4,5-bisphosphate (PIP2). It is thought that PIP2 stabilizes the activated state of the voltage sensor [52-54]. In contrast to cAMP, the effect of PIP2 is equal across the four HCN isoforms [53, 55]. Similar to the effect of PIP2, cholesterol has also a regulatory effect on the HCN channels [56, 57]. Cholesterol has been shown to decrease the localization of HCN4 into lipid rafts, leading to a redistribution of the channels within the membrane and modifying the kinetic properties. In addition, it shifted the midpoint of activation to more positive potentials and increased diastolic depolarization in rabbit SAN cells [56]. Fürst et al. showed that cholesterol uniquely regulates both channel expression and gating kinetics for each channel isoform [57]. Very recently, Liu et al. described a protein, cohesin-protein Shugosin-1 (SGO1), to play a crucial role in maintaining cardiac automaticity [58]. The function of SGO1 is to stabilize the cohesins that are important in the alignment of chromosomes during the G1 stage of mitosis. Next to its influence on cell division, it is shown that SGO1 colocalizes with HCN4. By regulating the HCN4 surface expression, it affects If, and thus, pacemaker activity [58]. Furthermore, channels can be activated by an independent mechanism, namely tyrosine phosphorylation by Src kinase. While HCN2 and HCN4 show an accelerated activation upon phosphorylation, phosphorylation does not affect HCN1 [59-62]. Additionally, the HCN channels are activated by bradykinin receptor stimulation, by the activation of phospholipase C, and by hormones released by the adrenergic and renin-angiotensin-aldosterone system (RAAS) [63]. In the study by Muto et al., they tested the effect of 10 nM aldosterone on cultured neonatal rat ventricular myocytes. The rate of spontaneous pulsing was found to be increased and, in the meantime, an upregulation of the mRNA and protein expression of HCN2 and HCN4 was observed. The latter effect was shown to be abolished by mineralocorticoid receptor (MR) antagonists (eplerenone and spironolactone), reducing the expression levels and currents of HCN2 and HCN4 to a level even lower than in control [64]. These findings were later confirmed by Song et al. [65]. Yu et al. further investigated the mechanism by which spironolactone regulates HCN channel expression after myocardial infarction (MI). Spironolactone increased miRNA-1 expression in rat ischemic left ventricular myocardium after MI. The upregulation of miRNA-1 (specifically expressed in skeletal muscle and cardiac myocytes) partially contributes to the posttranscriptional suppression of the HCN channel [66].

5. EXPRESSION AND PHYSIOLOGICAL ROLE

HCN channels play a central role in generating cardiac automaticity [67]. In general, HCN transcripts and proteins are expressed at the highest levels in the SAN and the conduction system, including an atrioventricular node (AVN) and Purkinje fibres. In the mammalian heart, the expression levels of individual HCN isoforms vary strongly in different cardiac regions. Since HCN3 shows a weak overall expression, the term “cardiac isoforms” generally refers to HCN1, HCN2, and HCN4 [68, 69]. Also, across different species, the expression levels of individual HCN isoforms differ. Comparable to the SAN of humans, the HCN4 isoform plays a crucial role in automaticity in rabbits, mice, and dogs. Additionally, the HCN1 and HCN2 isoforms can, albeit to a lesser extent, be found in the human heart. On the other hand, only the HCN1 isoform is present in mice and rabbits, and HCN2 can only be found in the SAN of the rat at levels similar to HCN4 [70]. The expression pattern of the HCN protein in the atria and the ventricles also shows isoform variability; however, most mammalians, including humans, exhibit a prevalence of HCN4 and HCN2, followed by a smaller amount of HCN1 and negligible levels of HCN3 [71, 72]. In the human AVN, there is a lack of HCN2 and low expression of HCN1. All three cardiac HCN isoforms have higher expression in the SAN than in the atria. Moreover, since it is almost exclusively present in SAN, HCN1 might be indicated as a selective molecular marker of human SAN and as a potential target of specific treatments intended to modify sinus rhythm [69].

The HCN channels have been recognized for their primary role in the generation and the regulation of pacemaker activity. The specific mechanism of HCN channels has considerably been elucidated in recent years. The role of HCN channels in the SAN is related to the generation and regulation of cardiac pacemaking. However, the relative contribution of the HCN-dependent voltage clock to cardiac automaticity towards the Ca2+ clock remains a matter of debate [73-75]. The two main mechanisms, the voltage clock and the Ca2+ clock, depend on the HCN current and the Ca2+ release from the sarcoplasmic reticulum, respectively. They seem to contribute to a coordinated system that supports spontaneous electrical activity in SAN [76]. In the AVN, the function of HCN channels does not differ substantially. Although the specific role played by HCN channels in this region is less investigated, previous studies have shown that HCN channels are implicated in AVN pacemaking and conduction [77-80]. Atrial and ventricular cardiomyocytes, on the other hand, do not usually display spontaneous activity. These cells have a stable resting membrane potential and infrequently display electrogenesis, consistent with a low contribution of the HCN current to resting membrane potentials (-80/-70 mV) and absence of spontaneous automaticity [63, 81]. A study by Fenske et al. showed a different function of HCN in the mouse ventricle. The HCN channel appears to contribute to the ventricular action potential (AP) waveform, specifically during late repolarization. This suggests that HCN channels, and particularly HCN3, provide a background conductance in cardiac myocytes that regulates the resting membrane potential and controls the kinetics of repolarization [82].

Mutations in ion channels and the knockout of ion channels in mice have been linked to sick sinus syndrome or bradycardia. Among these channels, HCN1-deficient mice display congenital SAN dysfunction characterized by bradycardia, and accompanied by low cardiac output, sinus dysrhythmia, and recurrent sinus pauses. This observation confirms the primary role of HCN1 in cardiac pacemaking [83]. Other studies have focused on the physiological role of HCN2. Mice lacking the HCN2 isoform showed pronounced cardiac sinoatrial dysrhythmia. These results reveal that the current generated by HCN2 channels critically determines the resting membrane potential of cardiac pacemaker cells. Nevertheless, the loss of HCN2 does not abolish the spontaneous activity of cardiac pacemaking cells, suggesting that multiple currents underlie pacemaking [84]. Furthermore, transgenic mice overexpressing HCN2 specifically in their hearts (HCN2-Tg) were analyzed. HCN2-Tg mice exhibited no noticeable abnormalities under physiological conditions, except for a significantly faster heart rate. Meanwhile, under pathological conditions, such as an excessive β-adrenergic stimulation, an enhancement in HCN2 channel activity was noticed, reducing the repolarization reserve of the ventricular AP and increasing the ectopic automaticity of ventricular myocytes. These results suggest that, in heart failure, increased HCN2 channel expression alone is not sufficient to induce lethal arrhythmia, but potentially increases the susceptibility to arrhythmias under sympathetic stimulation [85, 86]. Moreover, HCN2 overexpression also increases the vulnerability to arrhythmia under hypokalemic conditions, which is an electrolyte disturbance characterized by an abnormally low level of potassium [87]. Quantitatively, in all vertebrates studied so far, HCN4 underlies the major fraction of SAN current, accounting for 70% to 80% of the total If. Mice, deficient in the HCN4 isoform, were used to determine the physiological role of the respective channels. Stieber et al. observed mice lacking HCN4 channels, as well as mice with a selective deletion of HCN4 in cardiomyocytes, dying between embryonic days 9.5 and 11.5 While cardiac cells with primitive pacemaker action potentials were found both in wild-type (WT) and HCN4-deficient mice, a mature sinoatrial node-like pacemaker potential did only develop in wild-type embryos starting at day 9.0. This observation indicates that HCN4 is essential for the formation of mature pacemaker cells during embryogenesis [88]. Another study, conducted by the same research group, showed that deletion of HCN4 in adult mice reduced the sinoatrial If, on average by about 75%, and resulted in a cardiac arrhythmia characterized by recurrent sinus pauses. Surprisingly, they noticed that the channel does not play a role in the β-adrenergic-induced increase in heart rate, suggesting that HCN4 functions as a depolarization reserve supporting pacemaking in certain critical physiological states [89]. Transgenic mouse models in the study by Baruscotti et al. showed that the selective reduction of the HCN4 current is consistently accompanied by progressive development of severe bradycardia and AV block, eventually resulting in cardiac arrest and death of the animal. These contradicting results may be related to the inducible and cardiac-specific HCN4 knockout mouse model used [90]. Abolition of HCN4 sensitivity through the conditional expression of dominant-negative HCN4 channels lacking cAMP sensitivity reduces the spontaneous activity of AVN cells under basal conditions but does not impair the maximal response to β-adrenergic stimulation. This suggests that HCN4 channels influence basal excitability of mouse AV node cells, but are not required for acceleration of pacemaker activity under β-adrenergic receptor stimulation [78]. On the other hand, overexpression of the HCN4 isoform resulted in mice developing a dilated cardiomyopathy phenotype with increased cellular arrhythmogenicity, while the heart rate and conduction parameters remained unchanged. They did notice a change in the Na+/Ca2+ exchanger equilibrium. Due to the increased If, a diastolic Na+ influx causes an increased intracellular calcium concentration, which in turn creates significantly higher systolic intracellular calcium transients and stimulates apoptosis [91]. Therefore, HCN2 and HCN4 channels can be said to play a crucial role in the normal rhythmicity of sinoatrial myocytes.

6. PATHOLOGY

Atrial fibrillation (AF) is the most common persistent arrhythmia. A well-known risk factor is older age, with approximately 70% of AF patients being 65-85 years old [92]. Li et al. investigated the relationship between ageing and changes in HCN channel distribution in the sinoatrial node and atrium. In the SAN of the aged dogs, the expression level of HCN2 and HCN4 channel mRNAs and proteins was lower, while there was an upregulation in the atrium. Aged dogs presented a higher induction rate of AF in response to electrical stimulation, longer AF duration after induction, and longer sinus node recovery time compared to adults [93]. The age-associated expression of the HCN channel isoforms in rat SAN was also investigated by Huang et al. The protein levels of HCN2 and HCN4 decreased in 30-month-old rats compared to 4-month-old adult rats, corroborating the age-dependent lower intrinsic heart rate. Additionally, the effect of ivabradine, an HCN channel blocker, on rat SAN automaticity decreased [70]. HCN4 protein expression in aged rats is reduced, not only in the SAN, but also in the bundle of His within the AVN. These results indicate that HCN4 plays an important role in AVN dysfunction with ageing [94].

Another factor associated with an increased risk of arrhythmias is diabetes mellitus (DM). Huang et al. used streptozotocin (STZ)-induced diabetic rats to investigate the expression of the four HCN isoforms. The SAN of these animals showed a reduction in transcripts and proteins of HCN2 and HCN4 compared to the age-matched controls. The STZ-induced diabetic rats were characterized by cardiac alterations (lower intrinsic heart rate, lengthened sinoatrial conduction time, and rate-corrected maximal sinoatrial node recovery time in vivo as well as a longer cycle length in vitro) and by an impaired SAN function (inferior leading pacemaker site, reduced SAN conduction velocity and diastolic depolarization slope, and longer AP duration) [95].

In accordance with the physiological role of the HCN channels, genetic mutations modifying If functions are related to the genetic basis of arrhythmias. In humans, these channel mutations are exclusively limited to the HCN4 isoform and rarely to KCNE2 (Fig. 2). A review by Verkerk and Wilders reported 22 mutations or variants in HCN4 associated with sinus node dysfunction (SND) [96]. However, after evaluation of a clear genotype-phenotype association, this number reduced to 13 [97]. SND, also termed sick sinus syndrome (SSS), is defined as the “intrinsic inadequacy of the SAN to perform its pacemaking function due to a disorder of automaticity and/or inability to transmit its impulse to the rest of the atrium”. The heart rhythm can be too fast, too slow, interrupted by long pauses (>2 or 3 seconds), or an alternating combination of these problems [96]. Difrancesco et al. stated that all mutations are heterozygous and all mutations are dominant-negative, with various degrees of penetrance. In addition, they mentioned that all are loss-of-function (LOF) mutations. LOF is defined as a functional loss caused either by a negative shift of the activation curve and/or by the lower density of membrane expression of channels, and/or by a consequent reduction in current density and/or cAMP sensitivity variations [96, 97]. Möller et al. described two more homomeric channel mutants (R550H and E1193Q) showing LOF through increased rates of deactivation. A third channel mutation, R378C, exhibited a shift to the left in the activation curve and a slower activation rate. Possibly for the three mutants, this was caused by a significantly reduced functional HCN4 channel availability and cell surface expression due to defective trafficking [98]. The E1193Q mutant is located in the distal C-terminus and was previously described in a patient with paroxysmal AF [99]. Genetic screening of patients with suspected or diagnosed Brugada or SSS identified a novel mutation, V492F, located in S6, a highly conserved site of HCN4. Unlike other HCN4 mutants, this LOF can be fully attributed to a reduction in channel activity, while other mutations impact the synthesis and/or trafficking of the channel protein. Heteromeric channels, composed of WT and mutant monomers, show a shift in activation to more negative voltages, partially rescuing the If [100]. Among the SND, there is an inappropriate sinus tachycardia (IST). A familial form of IST is associated with a first gain-of-function (GOF) mutation (R524Q), contradicting the initial general observations by Difrancesco et al. Ex vivo analysis showed enhanced sensitivity to cAMP, leading to an increased pacemaker current during diastole [101, 102]. In addition to its known association with SND, HCN4 channel dysfunction may play a distinct role in the development of structural cardiac abnormalities. In family members with SND and noncompaction cardiomyopathy (NCCM), a new HCN4 LOF mutation G482R, located in the highly conserved channel pore domain, was identified and segregated with a combined disease phenotype. NCCM is characterized by excessive ventricular hypertrabeculation and complicated by heart failure, arrhythmia, and thromboembolic events. Moreover, due to a different truncation (695X) and a missense (P883R) HCN4 mutation segregated with a similar combined phenotype in an additional, unrelated family and a single unrelated proband, respectively, it can be concluded that HCN4 mutations are also critically involved in the development of NCCM [103]. Weigl et al. identified the previously uncharacterized HCN4 mutant, P883R, in three unrelated patients with AF and tachycardia-induced cardiomyopathy (TIC). The P883R mutation, however, was shown to be a new GOF mutation. Unlike the R524Q GOF mutation, P883R channels were characterized by a positive shift of the half-maximal activation voltage, while channel deactivation was faster. Moreover, net sensitivity towards cAMP was reduced. Co-transfection of WT and mutant channel, resembling the heterozygous cellular conditions of the patients, showed significantly higher current densities compared to WT. Altogether, these results could explain the increase in ectopic trigger and maintenance of AF [104].

Fig. (2).

Fig. (2)

Mutations in KCNE2 and the HCN4 isoform associated with sinus node dysfunction. Figure adapted from Servier Medical Art by Servier. https://smart.servier.com/(CreativeCommonsAttribution3.0UnportedLicense.)

As indicated earlier, the KCNE2 gene encoding the MiRP1 protein can carry mutations. Nawathe et al. identified a 55-year-old patient with the M54T MiRP1 mutation, presenting sinus bradycardia, and further investigated this mutation in human HCN2 and HCN4 isoforms. The voltage dependence of neither HCN2 nor HCN4 was altered by M54T. Additionally, while the current amplitude of HCN4 decreased, no effects were seen on the current amplitude of HCN2. Finally, M54T slows activation kinetics of HCN2 and HCN4 at physiologically relevant voltages but does not alter HCN4 deactivation kinetics [105]. There is evidence to support key roles of HCN-KCNE2 complexes in establishing susceptibility to SND. The expression patterns of KCNE2 in cardiac tissue resemble that of HCN channels, and gene variants alter the stoichiometry and increase the susceptibility to several of the same cardiac disorders linked to HCN, including cardiac arrhythmias [30].

7. PHARMACOLOGY

7.1. Current Controversies Concerning Ivabradine

Ivabradine, developed by Servier®, is the one molecule currently on the market acting as a specific HCN channel blocker. The European Medicines Agency (EMA) approved ivabradine for the treatment of stable angina and the management of patients with HF in 2005 and 2012, respectively. The Food and Drug Administration (FDA) only approved the drug to reduce rehospitalization rates in patients with chronic and stable HF in 2015 [8]. Ivabradine is licensed under different names like Procoralan®, Corlentor®, and Corlanor®, and is prescribed for patients with HF, with an ejection fraction ≤ 35% and a heart rate ≥ 70 bpm. In comparison to other treatments to control the heart rate (HR), such as β-blockers and calcium channel inhibitors, the effect of ivabradine is not associated with common side effects, like hypotension and a negative inotropic effect.

Bucchi et al. showed that HCN4 is inhibited by ivabradine in a use-dependent manner, with IC50 in the micromolar range. While ivabradine is an open-channel blocker of HCN4, it is a closed-channel blocker of HCN1 channels [106, 107]. Ivabradine exhibits a high selectivity for HCN channels, but various studies have shown that higher concentrations of this drug may also affect other cardiac ion channels [108]. A study by Delpon et al. showed a block of Kv1.5 by ivabradine, with an IC50 value of 29.0 ± 1.9 µM [109]. Also, Kv11.1 and Nav1.5 are inhibited by ivabradine at micromolar concentrations in heterologous systems [110]. T-type calcium currents are not affected by ivabradine (10 µM), while L-type calcium currents are decreased by <20% after the application of 3 µM of ivabradine. The use of ivabradine is restricted due to the lack of detailed information concerning subtype selectivity among HCN1–HCN4 channels, but it is thought to be a preferential HCN4 blocker [8]. Lack of specific selectivity may explain the side effects, such as bradycardia, hypertension, atrial fibrillation, and temporary visual disturbances, known as phosphenes [111].

7.2. HCN Channel as a Drug Target

7.2.1. Peptide Toxins

To date, only one peptide toxin is found to modulate HCN channels. Gamma-conotoxin-PnVIIA is described from the venom of the molluscivorous snail Conus pennaceus to belong to the six cysteines four-loop structural family of conotoxins [112]. Fainzilber et al. purified the toxin and characterized it by conducting electrophysiology experiments. First, caudodorsal neurons from the snail Lymnaea stagnalis were used to investigate the effects on action potential firing and membrane potential. It was shown that PnVIIA caused an increased excitability of these neuroendocrine cells and an increased number of action potentials, both in a dose-dependent manner. Besides the dose, the experiments showed that the duration of PnVIIA application influenced the response. The membrane potential was investigated by injecting hyperpolarizing current pulses into the cells, causing hyperpolarization of the membrane potential. Upon application of PnVIIA, the hyperpolarizing response decreased, indicating a decrease in the membrane resistance. They concluded that PnVIIA leads to the opening of ion channels. In the next series of experiments, they used the whole-cell voltage-clamp configuration to investigate the nature of these ion channels. Since no effect was seen on Nav or Cav channels, they believed PnVIIA modulates the pacemaker channels. After the addition of 10 µM PnVIIA, the activation curve shifted to more negative potentials by ~10 mV, and the non-inactivating outward current (at potentials above 0 mV) increased. However, further studies are required to elucidate the mechanism of action [112].

In a case study by Agarwal et al., they reported sinus node dysfunction following a snake bite for the first time [113]. Two hours after the snake bite, a 35-year-old man was admitted to the hospital. As seen in similar cases, the man presented with coagulopathy, an increased prothrombin time, and an increased activated partial thromboplastin time (aPTT). Nevertheless, he was hemodynamically stable, and after a complete check-up, his other parameters seemed normal. His electrocardiogram (ECG), on the other hand, displayed sinus arrest with junctional escape rhythm and retrograde P waves, at a rate of 40 beats/min. After 20 vials of anti-snake venom and 3 days of hospital admission, the ECG showed normal sinus rhythm. They ruled out the diagnosis of toxic myocarditis and suggested that the snake venom caused a modification in the electrophysiological properties of the cardiac cell membrane, and hence, affected impulse generation and conduction. This venom could be explored as a possible source of novel peptide toxins modulating the HCN channel. However, they mentioned that future research is also required [113].

7.2.2. Small Molecules

To date, several pharmacological agents have been shown to exert an effect on the heart rate. An extensive and thorough review of these drugs has been published by Romanelli et al. [108]. They summarized the current knowledge regarding the compounds that have been shown to modulate HCN channels, including α2-adrenergic and opioid receptor agonists (dexmedetomidine, clonidine, loperamide, tramadol), general anaesthetics (propofol, ketamine), local anaesthetics, antiarrhythmic drugs (lidocaine, mepivacaine, bupivacaine, dronedarone), and several others [108, 114-116].

There is a great need for novel potent and selective HCN channel inhibitors as the use of ivabradine is limited. Compounds selective for HCN1 have been described by McClure et al. One of the compounds is 15 times more selective for HCN1 than for HCN4 [117]. Carvedilol is a third-generation β1-blocker, clinically effective in the treatment of chronic HF and exhibiting favourable effects compared to other β-blockers. HCN4 current significantly reduces when carvedilol is used in the low micromolar range. In addition, these effects are also shown for HCN1 and HCN2, but with a higher IC50. The rate of channel activation is found to decrease, that of deactivation to increase, and the voltage-dependence of activation to shift to more hyperpolarizing potentials [118]. Chen et al. developed and synthesized a series of alkanol amine derivatives [119]. The compound 4e blocked the HCN2 current with an IC50 value of 2.9 ± 1.2 μM. Its application caused a slowing of activation and a shift to more negative potentials in the voltage dependence of HCN2 channel activation. Attempts at achieving a selective inhibition of individual HCN isoforms have been performed through the design of a series of phenylalkylamines related to zatebradine, which is a known blocker of If in cardiac cells [120]. The compounds EC18 and MEL57A showed preferential block for HCN4 and HCN1, respectively. They tested selectivity on HCN channels expressed in HEK cells. Subsequent tests showed that the effect was preserved in native tissues, where the subunits are thought to form heteromeric channels. In guinea pig SAN cells expressing HCN4 channels, only EC18 displayed an effect, while in DRG neurons, where HCN1 is the most relevant contributor to native Ih, only MEL57A exhibited a significant effect. In dog Purkinje fibres, where HCN4 is the prevalent isoform, only EC18 reduced the amplitude and decreased the slope of the diastolic depolarization phase [121-123]. However, since it is thought that HCN1 and HCN2 channels can co-assemble and form heterotetramers, MEL55A shows an interesting pharmacological profile, exhibiting a preferential block for HCN1 and HCN2 over HCN4 [122, 124]. β-blockers have been shown to reduce morbidity and mortality in patients with HF, but the effect on the SAN function is still unclear. Du et al. presented a study where bisoprolol improved SAN function in rats with HF by reversing the down-regulation of the HCN4 channel [125]. Also, the mechanism of ethanol on the SAN is still not completely understood. Romanelli et al. previously mentioned that the application of ethanol shifts the voltage dependence of HCN channels to more depolarized potentials and increases the maximum conductance, causing an increase in If, and thus, in spontaneous firing [108, 126]. These results were confirmed by Chen et al. [127]. Recently, the first high-throughput screening of 16000 small-molecule compounds using an automated patch-clamp system was conducted to identify novel HCN4 blockers [128]. Three novel hit compounds were identified, T-478, T-788 and T-524, with IC50 values of 4.2 µM, 0.97 µM and 10.3 µM, respectively. All three compounds exhibited different blockade curves. While T-478 showed similar curves to ivabradine, the curves of T-788 and T-524 were similar to ZD7288. Finally, two compounds, already used for several years as herbal medicine, have been studied on HCN channels. A recent paper by Chan et al. suggested honokiol (HNK) as a potential modulator of the HCN channel [129]. HNK is a small-molecule polyphenol obtained from Magnolia officinalis, and has been used in traditional Asian medicines. They investigated the effect of HNK on the density and gating of Ih in pituitary tumor (GH3) cells and in Rolf B1.T olfactory neurons. The density of Ih was found to be decreased and the activation curve shifted to more hyperpolarized potentials. The IC50 value was estimated to be 2.1 μM. Further experiments are required to investigate the effect of HNK on If in cardiac myocytes, where it could possibly reduce the rhythmic activity of the heart. Next, the effect of Ginkgo Biloba Extract (GBE) had been tested on HCN channels [130]. It has been shown previously that GBE decreased the slope of the diastole and inhibited If in SAN cells [131, 132]. HCN2 and HCN4 channel currents are inhibited by GBE in a concentration-dependent manner. HCN4 was shown to be more sensitive than HCN2, with IC50 values of 0.12 ± 0.05 mg/ml and 0.25 ± 0.01 mg/ml, respectively. This antiarrhythmic mechanism of GBE might be useful for the treatment of arrhythmia.

CONCLUSION AND FUTURE PERSPECTIVE

The evidence that HCN channels are important drug targets in the therapy of several cardiovascular diseases is overwhelming. Dysfunction of HCN channels plays a crucial role in cardiac arrhythmias, ischemic heart disease, and ventricular hypertrophy [133]. Due to their unique features, HCN channels are present in pacemaker cells only, and are emerging as ideal targets for the modulation of pacemaking. The current therapies are rather limited. For years, research has been seeking compounds for lowering the heart rate without the side effects of β-blockers and calcium antagonists. Hence, these HCN channels are seen as a valuable alternative for the development of new clinical therapies. Although ivabradine is clinically approved as a treatment for stable angina pectoris and is a great tool to investigate the mechanism of the HCN channels, it is bound by a major limitation, i.e., it is equipotent across the different HCN channel isoforms. Since non-selective modulation of HCN channels can cause unwanted side effects, future discoveries of novel and HCN channel isoform-selective compounds are required. The currently available substances, including ivabradine, inhibit HCN channels through the same mechanism; they block the channel pore [106, 134]. Since the pore region is extremely conserved across the channel isoforms, the chances of discovering isoform-selective molecules may be higher when targeting other regions of the channel. The recently published cryo-electron microscopy (EM) structure of HCN1, by Lee and MacKinnon, can lay out a framework for structure-guided drug discovery [12]. HCN channel isoform-selective drug design will be facilitated when the cryo-EM structures of HCN2 and HCN4 will be discovered, and subsequently, the differences in the channel structures of the different isoforms could be determined. Furthermore, the cryo-EM structure of HCN1 will provide the means for understanding the mechanism of action of HCN channels.

Future research is required to discover isoform-selective molecules modulating HCN channels, and to further determine the physiological role of HCN channels. This knowledge will be fundamental for the treatment of a variety of HCN channelopathies.

ACKNOWLEDGEMENTS

Declared none.

CONSENT FOR PUBLICATION

Not applicable.

FUNDING

This study was supported by grants G0E7120N, GOC2319N, and GOA4919N from the F.W.O. Vlaanderen awarded to J.T., and S.P. was supported by KU Leuven funding (PDM/19/164) and F.W.O. Vlaanderen grant 12W7822N.

CONFLICT OF INTEREST

The authors declare no conflict of interest, financial or otherwise.

REFERENCES

  • 1.Bennett J.E., Stevens G.A., Mathers C.D., et al. NCD Countdown 2030 collaborators. NCD Countdown 2030: worldwide trends in non-communicable disease mortality and progress towards Sustainable Development Goal target 3.4. Lancet. 2018;392(10152):1072–1088. doi: 10.1016/S0140-6736(18)31992-5. [DOI] [PubMed] [Google Scholar]
  • 2.Wahl-Schott C., Fenske S., Biel M. HCN channels: new roles in sinoatrial node function. Curr. Opin. Pharmacol. 2014;15:83–90. doi: 10.1016/j.coph.2013.12.005. [DOI] [PubMed] [Google Scholar]
  • 3.Rivolta I., Binda A., Masi A., DiFrancesco J.C. Cardiac and neuronal HCN channelopathies. Pflugers Arch. 2020;472(7):931–951. doi: 10.1007/s00424-020-02384-3. [DOI] [PubMed] [Google Scholar]
  • 4.Alig J., Marger L., Mesirca P., Ehmke H., Mangoni M.E., Isbrandt D. Control of heart rate by cAMP sensitivity of HCN chan-nels. Proc. Natl. Acad. Sci. USA. 2009;106(29):12189–12194. doi: 10.1073/pnas.0810332106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Wahl-Schott C., Biel M. HCN channels: structure, cellular regulation and physiological function. Cell. Mol. Life Sci. 2009;66(3):470–494. doi: 10.1007/s00018-008-8525-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Postea O., Biel M. Exploring HCN channels as novel drug targets. Nat. Rev. Drug Discov. 2011;10(12):903–914. doi: 10.1038/nrd3576. [DOI] [PubMed] [Google Scholar]
  • 7.DiFrancesco D., Borer J.S. The funny current: cellular basis for the control of heart rate. Drugs. 2007;67(Suppl. 2):15–24. doi: 10.2165/00003495-200767002-00003. [DOI] [PubMed] [Google Scholar]
  • 8.Chaudhary R., Garg J., Krishnamoorthy P., et al. Ivabradine: heart failure and beyond. J. Cardiovasc. Pharmacol. Ther. 2016;21(4):335–343. doi: 10.1177/1074248415624157. [DOI] [PubMed] [Google Scholar]
  • 9.Liman E.R., Tytgat J., Hess P. Subunit stoichiometry of a mammalian K+ channel determined by construction of mul-timeric cDNAs. Neuron. 1992;9(5):861–871. doi: 10.1016/0896-6273(92)90239-A. [DOI] [PubMed] [Google Scholar]
  • 10.Wang H., Kunkel D.D., Martin T.M., Schwartzkroin P.A., Tempel B.L., Heteromultimeric K. Heteromultimeric K+ channels in terminal and juxtaparanodal regions of neurons. Nature. 1993;365(6441):75–79. doi: 10.1038/365075a0. [DOI] [PubMed] [Google Scholar]
  • 11.Sheng M., Liao Y.J., Jan Y.N., Jan L.Y. Presynaptic A-current based on heteromultimeric K+ channels detected in vivo. Nature. 1993;365(6441):72–75. doi: 10.1038/365072a0. [DOI] [PubMed] [Google Scholar]
  • 12.Lee C.H., MacKinnon R. Structures of the human HCN1 hy-perpolarization-activated channel. Cell. 2017;168(1-2):111–120.e11. doi: 10.1016/j.cell.2016.12.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Ulens C., Tytgat J. Functional heteromerization of HCN1 and HCN2 pacemaker channels. J. Biol. Chem. 2001;276(9):6069–6072. doi: 10.1074/jbc.C000738200. [DOI] [PubMed] [Google Scholar]
  • 14.Zagotta W.N., Olivier N.B., Black K.D., Young E.C., Olson R., Gouaux E. Structural basis for modulation and agonist speci-ficity of HCN pacemaker channels. Nature. 2003;425(6954):200–205. doi: 10.1038/nature01922. [DOI] [PubMed] [Google Scholar]
  • 15.Gauss R., Seifert R., Kaupp U.B. Molecular identification of a hyperpolarization-activated channel in sea urchin sperm. Nature. 1998;393(6685):583–587. doi: 10.1038/31248. [DOI] [PubMed] [Google Scholar]
  • 16.Santoro B., Liu D.T., Yao H., et al. Identification of a gene en-coding a hyperpolarization-activated pacemaker channel of brain. Cell. 1998;93(5):717–729. doi: 10.1016/S0092-8674(00)81434-8. [DOI] [PubMed] [Google Scholar]
  • 17.Ludwig A., Zong X., Jeglitsch M., Hofmann F., Biel M. A family of hyperpolarization-activated mammalian cation channels. Nature. 1998;393(6685):587–591. doi: 10.1038/31255. [DOI] [PubMed] [Google Scholar]
  • 18.Whicher J.R., MacKinnon R. Structure of the voltage-gated K channel Eag1 reveals an alternative voltage sensing mecha-nism. Science. 2016;353(6300):664–669. doi: 10.1126/science.aaf8070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Long S.B., Campbell E.B., Mackinnon R. Voltage sensor of Kv1.2: structural basis of electromechanical coupling. Science. 2005;309(5736):903–908. doi: 10.1126/science.1116270. [DOI] [PubMed] [Google Scholar]
  • 20.Long S.B., Tao X., Campbell E.B., MacKinnon R. Atomic struc-ture of a voltage-dependent K+ channel in a lipid membrane-like environment. Nature. 2007;450(7168):376–382. doi: 10.1038/nature06265. [DOI] [PubMed] [Google Scholar]
  • 21.Chen J., Mitcheson J.S., Lin M., Sanguinetti M.C. Functional roles of charged residues in the putative voltage sensor of the HCN2 pacemaker channel. J. Biol. Chem. 2000;275(46):36465–36471. doi: 10.1074/jbc.M007034200. [DOI] [PubMed] [Google Scholar]
  • 22.Bezanilla F. The voltage sensor in voltage-dependent ion channels. Physiol. Rev. 2000;80(2):555–592. doi: 10.1152/physrev.2000.80.2.555. [DOI] [PubMed] [Google Scholar]
  • 23.Bell D.C., Yao H., Saenger R.C., Riley J.H., Siegelbaum S.A. Changes in local S4 environment provide a voltage-sensing mechanism for mammalian hyperpolarization-activated HCN channels. J. Gen. Physiol. 2004;123(1):5–19. doi: 10.1085/jgp.200308918. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Cowgill J., Klenchin V.A., Alvarez-Baron C., Tewari D., Blair A., Chanda B. Bipolar switching by HCN voltage sensor underlies hyperpolarization activation. Proc. Natl. Acad. Sci. USA. 2019;116(2):670–678. doi: 10.1073/pnas.1816724116. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Ramentol R., Perez M.E., Larsson H.P. Gating mechanism of hyperpolarization-activated HCN pacemaker channels. Nat. Commun. 2020;11(1):1419. doi: 10.1038/s41467-020-15233-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Flynn G.E., Zagotta W.N. Insights into the molecular mecha-nism for hyperpolarization-dependent activation of HCN channels. Proc. Natl. Acad. Sci. USA. 2018;115(34):E8086–E8095. doi: 10.1073/pnas.1805596115. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.Brandt M.C., Endres-Becker J., Zagidullin N., et al. Effects of KCNE2 on HCN isoforms: distinct modulation of membrane expression and single channel properties. Am. J. Physiol. Heart Circ. Physiol. 2009;297(1):H355–H363. doi: 10.1152/ajpheart.00154.2009. [DOI] [PubMed] [Google Scholar]
  • 28.Qu J., Kryukova Y., Potapova I.A., et al. MiRP1 modulates HCN2 channel expression and gating in cardiac myocytes. J. Biol. Chem. 2004;279(42):43497–43502. doi: 10.1074/jbc.M405018200. [DOI] [PubMed] [Google Scholar]
  • 29.Yu H., Wu J., Potapova I., et al. MinK-related peptide 1: A beta subunit for the HCN ion channel subunit family enhances ex-pression and speeds activation. Circ. Res. 2001;88(12):E84–E87. doi: 10.1161/hh1201.093511. [DOI] [PubMed] [Google Scholar]
  • 30.Lussier Y., Fürst O., Fortea E., et al. Disease-linked mutations alter the stoichiometries of HCN-KCNE2 complexes. Sci. Rep. 2019;9(1):9113. doi: 10.1038/s41598-019-45592-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Barbuti A., Scavone A., Mazzocchi N., Terragni B., Baruscotti M., Difrancesco D. A caveolin-binding domain in the HCN4 channels mediates functional interaction with caveolin pro-teins. J. Mol. Cell. Cardiol. 2012;53(2):187–195. doi: 10.1016/j.yjmcc.2012.05.013. [DOI] [PubMed] [Google Scholar]
  • 32.Ye B., Balijepalli R.C., Foell J.D., et al. Caveolin-3 associates with and affects the function of hyperpolarization-activated cyclic nucleotide-gated channel 4. Biochemistry. 2008;47(47):12312–12318. doi: 10.1021/bi8009295. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Haitin Y. Structural biology: a ‘funny’ cyclic dinucleotide receptor. Nat. Chem. Biol. 2014;10(6):413–414. doi: 10.1038/nchembio.1530. [DOI] [PubMed] [Google Scholar]
  • 34.Gross C., Saponaro A., Santoro B., Moroni A., Thiel G., Hamacher K. Mechanical transduction of cytoplasmic-to-transmembrane-domain movements in a hyperpolarization-activated cyclic nucleotide-gated cation channel. J. Biol. Chem. 2018;293(33):12908–12918. doi: 10.1074/jbc.RA118.002139. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35.Neyton J., Miller C. Potassium blocks barium permeation through a calcium-activated potassium channel. J. Gen. Physiol. 1988;92(5):549–567. doi: 10.1085/jgp.92.5.549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Macri V., Angoli D., Accili E.A. Architecture of the HCN selec-tivity filter and control of cation permeation. Sci. Rep. 2012;2(1):894. doi: 10.1038/srep00894. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Zhou Y., Morais-Cabral J.H., Kaufman A., MacKinnon R. Chemistry of ion coordination and hydration revealed by a K+ channel-Fab complex at 2.0 A resolution. Nature. 2001;414(6859):43–48. doi: 10.1038/35102009. [DOI] [PubMed] [Google Scholar]
  • 38.DiFrancesco D. A study of the ionic nature of the pace-maker current in calf Purkinje fibres. J. Physiol. 1981;314(1):377–393. doi: 10.1113/jphysiol.1981.sp013714. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.DiFrancesco D., Tortora P. Direct activation of cardiac pace-maker channels by intracellular cyclic AMP. Nature. 1991;351(6322):145–147. doi: 10.1038/351145a0. [DOI] [PubMed] [Google Scholar]
  • 40.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(6839):805–810. doi: 10.1038/35081088. [DOI] [PubMed] [Google Scholar]
  • 41.Wang J., Chen S., Siegelbaum S.A. Regulation of hyperpolariza-tion-activated HCN channel gating and cAMP modulation due to interactions of COOH terminus and core transmembrane regions. J. Gen. Physiol. 2001;118(3):237–250. doi: 10.1085/jgp.118.3.237. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.Mistrík P., Mader R., Michalakis S., Weidinger M., Pfeifer A., Biel M. The murine HCN3 gene encodes a hyperpolarization-activated cation channel with slow kinetics and unique re-sponse to cyclic nucleotides. J. Biol. Chem. 2005;280(29):27056–27061. doi: 10.1074/jbc.M502696200. [DOI] [PubMed] [Google Scholar]
  • 43.Stieber J., Stöckl G., Herrmann S., Hassfurth B., Hofmann F. Functional expression of the human HCN3 channel. J. Biol. Chem. 2005;280(41):34635–34643. doi: 10.1074/jbc.M502508200. [DOI] [PubMed] [Google Scholar]
  • 44.Ulens C., Siegelbaum S.A. Regulation of hyperpolarization-activated HCN channels by cAMP through a gating switch in binding domain symmetry. Neuron. 2003;40(5):959–970. doi: 10.1016/S0896-6273(03)00753-0. [DOI] [PubMed] [Google Scholar]
  • 45.Zong C., Lu S., Chapman A.R., Xie X.S. Genome-wide detection of single-nucleotide and copy-number variations of a single human cell. Science. 2012;338(6114):1622–1626. doi: 10.1126/science.1229164. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Ng L.C.T., Putrenko I., Baronas V., Van Petegem F., Accili E.A. Cyclic purine and pyrimidine nucleotides bind to the HCN2 ion channel and variably promote C-terminal domain interac-tions and opening. Structure. 2016;24(10):1629–1642. doi: 10.1016/j.str.2016.06.024. [DOI] [PubMed] [Google Scholar]
  • 47.Lolicato M., Bucchi A., Arrigoni C., et al. Cyclic dinucleotides bind the C-linker of HCN4 to control channel cAMP respon-siveness. Nat. Chem. Biol. 2014;10(6):457–462. doi: 10.1038/nchembio.1521. [DOI] [PubMed] [Google Scholar]
  • 48.DiFrancesco D. Cardiac pacemaker I(f) current and its inhibi-tion by heart rate-reducing agents. Curr. Med. Res. Opin. 2005;21(7):1115–1122. doi: 10.1185/030079905X50543. [DOI] [PubMed] [Google Scholar]
  • 49.DiFrancesco D. Serious workings of the funny current. Prog. Biophys. Mol. Biol. 2006;90(1-3):13–25. doi: 10.1016/j.pbiomolbio.2005.05.001. [DOI] [PubMed] [Google Scholar]
  • 50.Spinelli V., Sartiani L., Mugelli A., Romanelli M.N., Cerbai E. Hyperpolarization-activated cyclic-nucleotide-gated channels: pathophysiological, developmental, and pharmacological in-sights into their function in cellular excitability. Can. J. Physiol. Pharmacol. 2018;96(10):977–984. doi: 10.1139/cjpp-2018-0115. [DOI] [PubMed] [Google Scholar]
  • 51.VanSchouwen B., Melacini G. Regulation of HCN ion chan-nels by non-canonical cyclic nucleotides. Handb. Exp. Pharmacol. 2017;238:123–133. doi: 10.1007/164_2016_5006. [DOI] [PubMed] [Google Scholar]
  • 52.Pian P., Bucchi A., Robinson R.B., Siegelbaum S.A. Regulation of gating and rundown of HCN hyperpolarization-activated channels by exogenous and endogenous PIP2. J. Gen. Physiol. 2006;128(5):593–604. doi: 10.1085/jgp.200609648. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53.Zolles G., Klöcker N., Wenzel D., et al. Pacemaking by HCN channels requires interaction with phosphoinositides. Neuron. 2006;52(6):1027–1036. doi: 10.1016/j.neuron.2006.12.005. [DOI] [PubMed] [Google Scholar]
  • 54.Herrmann S., Schnorr S., Ludwig A. HCN channels--modulators of cardiac and neuronal excitability. Int. J. Mol. Sci. 2015;16(1):1429–1447. doi: 10.3390/ijms16011429. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.Ying S-W., Tibbs G.R., Picollo A., et al. PIP2-mediated HCN3 channel gating is crucial for rhythmic burst firing in thalamic intergeniculate leaflet neurons. J. Neurosci. 2011;31(28):10412–10423. doi: 10.1523/JNEUROSCI.0021-11.2011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Barbuti A., Gravante B., Riolfo M., Milanesi R., Terragni B., DiFrancesco D. Localization of pacemaker channels in lipid rafts regulates channel kinetics. Circ. Res. 2004;94(10):1325–1331. doi: 10.1161/01.RES.0000127621.54132.AE. [DOI] [PubMed] [Google Scholar]
  • 57.Fürst O., D’Avanzo N. Isoform dependent regulation of hu-man HCN channels by cholesterol. Sci. Rep. 2015;5:14270. doi: 10.1038/srep14270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Liu D., Song A.T., Qi X., et al. Cohesin-protein Shugoshin-1 controls cardiac automaticity via HCN4 pacemaker channel. Nat. Commun. 2021;12(1):2551. doi: 10.1038/s41467-021-22737-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Li C.H., Zhang Q., Teng B., Mustafa S.J., Huang J.Y., Yu H.G. Src tyrosine kinase alters gating of hyperpolarization-activated HCN4 pacemaker channel through Tyr531. Am. J. Physiol. Cell Physiol. 2008;294(1):C355–C362. doi: 10.1152/ajpcell.00236.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Yu H-G., Lu Z., Pan Z., Cohen I.S. Tyrosine kinase inhibition differentially regulates heterologously expressed HCN chan-nels. Pflugers Arch. 2004;447(4):392–400. doi: 10.1007/s00424-003-1204-y. [DOI] [PubMed] [Google Scholar]
  • 61.Zong X., Eckert C., Yuan H., et al. A novel mechanism of modulation of hyperpolarization-activated cyclic nucleotide-gated channels by Src kinase. J. Biol. Chem. 2005;280(40):34224–34232. doi: 10.1074/jbc.M506544200. [DOI] [PubMed] [Google Scholar]
  • 62.Arinsburg S.S., Cohen I.S., Yu H-G. Constitutively active Src tyrosine kinase changes gating of HCN4 channels through di-rect binding to the channel proteins. J. Cardiovasc. Pharmacol. 2006;47(4):578–586. doi: 10.1097/01.fjc.0000211740.47960.8b. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 63.Sartiani L., Romanelli M.N., Mugelli A., Cerbai E. Updates on HCN channels in the heart: function, dysfunction and phar-macology. Curr. Drug Targets. 2015;16(8):868–876. doi: 10.2174/1389450116666150531152047. [DOI] [PubMed] [Google Scholar]
  • 64.Muto T., Ueda N., Opthof T., et al. Aldosterone modulates I(f) current through gene expression in cultured neonatal rat ven-tricular myocytes. Am. J. Physiol. Heart Circ. Physiol. 2007;293(5):H2710–H2718. doi: 10.1152/ajpheart.01399.2006. [DOI] [PubMed] [Google Scholar]
  • 65.Song T., Yang J., Yao Y., et al. Spironolactone diminishes spontaneous ventricular premature beats by reducing HCN4 protein expression in rats with myocardial infarction. Mol. Med. Rep. 2011;4(3):569–573. doi: 10.3892/mmr.2011.455. [DOI] [PubMed] [Google Scholar]
  • 66.Yu H.D., Xia S., Zha C.Q., Deng S.B., Du J.L., She Q. Spironolac-tone regulates HCN protein expression through micro-RNA-1 in rats with myocardial infarction. J. Cardiovasc. Pharmacol. 2015;65(6):587–592. doi: 10.1097/FJC.0000000000000227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 67.Benarroch E.E. HCN channels: function and clinical implica-tions. Neurology. 2013;80(3):304–310. doi: 10.1212/WNL.0b013e31827dec42. [DOI] [PubMed] [Google Scholar]
  • 68.Herrmann S., Hofmann F., Stieber J., Ludwig A. HCN channels in the heart: lessons from mouse mutants. Br. J. Pharmacol. 2012;166(2):501–509. doi: 10.1111/j.1476-5381.2011.01798.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 69.Li N., Csepe T.A., Hansen B.J., et al. Molecular mapping of sinoatrial node HCN channel expression in the human heart. Circ. Arrhythm. Electrophysiol. 2015;8(5):1219–1227. doi: 10.1161/CIRCEP.115.003070. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 70.Huang X., Yang P., Yang Z., Zhang H., Ma A. Age-associated expression of HCN channel isoforms in rat sinoatrial node. Exp. Biol. Med. (Maywood) 2016;241(3):331–339. doi: 10.1177/1535370215603515. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 71.Stillitano F., Lonardo G., Zicha S., et al. Molecular basis of funny current (If) in normal and failing human heart. J. Mol. Cell. Cardiol. 2008;45(2):289–299. doi: 10.1016/j.yjmcc.2008.04.013. [DOI] [PubMed] [Google Scholar]
  • 72.Lezoualc’h F., Steplewski K., Sartiani L., Mugelli A., Fischmeis-ter R., Bril A. Quantitative mRNA analysis of serotonin 5-HT4 receptor isoforms, calcium handling proteins and ion chan-nels in human atrial fibrillation. Biochem. Biophys. Res. Commun. 2007;357(1):218–224. doi: 10.1016/j.bbrc.2007.03.124. [DOI] [PubMed] [Google Scholar]
  • 73.DiFrancesco D., Noble D. The funny current has a major pacemaking role in the sinus node. Heart Rhythm. 2012;9(2):299–301. doi: 10.1016/j.hrthm.2011.09.021. [DOI] [PubMed] [Google Scholar]
  • 74.Maltsev V.A., Lakatta E.G. The funny current in the context of the coupled-clock pacemaker cell system. Heart Rhythm. 2012;9(2):302–307. doi: 10.1016/j.hrthm.2011.09.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 75.Rosen M.R., Nargeot J., Salama G. The case for the funny cur-rent and the calcium clock. Heart Rhythm. 2012;9(4):616–618. doi: 10.1016/j.hrthm.2011.10.008. [DOI] [PubMed] [Google Scholar]
  • 76.Lakatta E.G., DiFrancesco D. What keeps us ticking: a funny current, a calcium clock, or both? J. Mol. Cell. Cardiol. 2009;47(2):157–170. doi: 10.1016/j.yjmcc.2009.03.022. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 77.Liu J., Noble P.J., Xiao G., et al. Role of pacemaking current in cardiac nodes: insights from a comparative study of sinoatrial node and atrioventricular node. Prog. Biophys. Mol. Biol. 2008;96(1-3):294–304. doi: 10.1016/j.pbiomolbio.2007.07.009. [DOI] [PubMed] [Google Scholar]
  • 78.Marger L., Mesirca P., Alig J., et al. Functional roles of Ca(v)1.3, Ca(v)3.1 and HCN channels in automaticity of mouse atrioventricular cells: insights into the atrioventricular pacemaker mechanism. Channels (Austin) 2011;5(3):251–261. doi: 10.4161/chan.5.3.15266. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 79.Verrier R.L., Bonatti R., Silva A.F.G., et al. If inhibition in the atrioventricular node by ivabradine causes rate-dependent slowing of conduction and reduces ventricular rate during atrial fibrillation. Heart Rhythm. 2014;11(12):2288–2296. doi: 10.1016/j.hrthm.2014.08.007. [DOI] [PubMed] [Google Scholar]
  • 80.Verrier R.L., Silva A.F.G., Bonatti R., et al. Combined actions of ivabradine and ranolazine reduce ventricular rate during atrial fibrillation. J. Cardiovasc. Electrophysiol. 2015;26(3):329–335. doi: 10.1111/jce.12569. [DOI] [PubMed] [Google Scholar]
  • 81.Cerbai E., Mugelli A.I. (f) in non-pacemaker cells: role and pharmacological implications. Pharmacol. Res. 2006;53(5):416–423. doi: 10.1016/j.phrs.2006.03.015. [DOI] [PubMed] [Google Scholar]
  • 82.Fenske S., Mader R., Scharr A., et al. HCN3 contributes to the ventricular action potential waveform in the murine heart. Circ. Res. 2011;109(9):1015–1023. doi: 10.1161/CIRCRESAHA.111.246173. [DOI] [PubMed] [Google Scholar]
  • 83.Fenske S., Krause S.C., Hassan S.I.H., et al. Sick sinus syndrome in HCN1-deficient mice. Circulation. 2013;128(24):2585–2594. doi: 10.1161/CIRCULATIONAHA.113.003712. [DOI] [PubMed] [Google Scholar]
  • 84.Ludwig A., Budde T., Stieber J., et al. Absence epilepsy and sinus dysrhythmia in mice lacking the pacemaker channel HCN2. EMBO J. 2003;22(2):216–224. doi: 10.1093/emboj/cdg032. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 85.Oshita K., Itoh M., Hirashima S., et al. Ectopic automaticity induced in ventricular myocytes by transgenic overexpres-sion of HCN2. J. Mol. Cell. Cardiol. 2015;80:81–89. doi: 10.1016/j.yjmcc.2014.12.019. [DOI] [PubMed] [Google Scholar]
  • 86.Kuwabara Y., Kuwahara K., Takano M., et al. Increased ex-pression of HCN channels in the ventricular myocardium contributes to enhanced arrhythmicity in mouse failing hearts. J. Am. Heart Assoc. 2013;2(3):e000150–e0. doi: 10.1161/JAHA.113.000150. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 87.Oshita K., Kozasa Y., Nakagawa Y., et al. Overexpression of the HCN2 channel increases the arrhythmogenicity induced by hypokalemia. J. Physiol. Sci. 2019;69(4):653–660. doi: 10.1007/s12576-019-00684-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 88.Stieber J., Herrmann S., Feil S., et al. The hyperpolarization-activated channel HCN4 is required for the generation of pacemaker action potentials in the embryonic heart. Proc. Natl. Acad. Sci. USA. 2003;100(25):15235–15240. doi: 10.1073/pnas.2434235100. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 89.Herrmann S., Stieber J., Stöckl G., Hofmann F., Ludwig A. HCN4 provides a ‘depolarization reserve’ and is not required for heart rate acceleration in mice. EMBO J. 2007;26(21):4423–4432. doi: 10.1038/sj.emboj.7601868. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 90.Baruscotti M., Bucchi A., Viscomi C., et al. Deep bradycardia and heart block caused by inducible cardiac-specific knock-out of the pacemaker channel gene Hcn4. Proc. Natl. Acad. Sci. USA. 2011;108(4):1705–1710. doi: 10.1073/pnas.1010122108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 91.Yampolsky P., Koenen M., Mosqueira M., et al. Augmentation of myocardial If dysregulates calcium homeostasis and causes adverse cardiac remodeling. Nat. Commun. 2019;10(1):3295. doi: 10.1038/s41467-019-11261-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 92.Fuster V., Rydé N.L.E., Cannom D.S. ACC/AHA/ESC 2006 Guidelines for the management of patients with atrial fibrilla-tion. Circulation. 2006;114(7):e257–e354. doi: 10.1161/CIRCULATIONAHA.106.177292. [DOI] [PubMed] [Google Scholar]
  • 93.Li Y.D., Hong Y.F., Zhang Y., et al. Association between rever-sal in the expression of hyperpolarization-activated cyclic nu-cleotide-gated (HCN) channel and age-related atrial fibrilla-tion. Med. Sci. Monit. 2014;20:2292–2297. doi: 10.12659/MSM.892505. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 94.Saeed Y., Temple I.P., Borbas Z., et al. Structural and functional remodeling of the atrioventricular node with aging in rats: The role of hyperpolarization-activated cyclic nucleotide-gated and ryanodine 2 channels. Heart Rhythm. 2018;15(5):752–760. doi: 10.1016/j.hrthm.2017.12.027. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 95.Huang X., Zhong N., Zhang H., Ma A., Yuan Z., Guo N. Re-duced expression of HCN channels in the sinoatrial node of streptozotocin-induced diabetic rats. Can. J. Physiol. Pharmacol. 2017;95(5):586–594. doi: 10.1139/cjpp-2016-0418. [DOI] [PubMed] [Google Scholar]
  • 96.Verkerk A.O., Wilders R. Pacemaker activity of the human sinoatrial node: an update on the effects of mutations in HCN4 on the hyperpolarization-activated current. Int. J. Mol. Sci. 2015;16(2):3071–3094. doi: 10.3390/ijms16023071. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 97.DiFrancesco D. HCN4, sinus bradycardia and atrial fibrilla-tion. Arrhythm. Electrophysiol. Rev. 2015;4(1):9–13. doi: 10.15420/aer.2015.4.1.9. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 98.Möller M., Silbernagel N., Wrobel E., et al. in vitro analyses of novel HCN4 gene mutations. Cell. Physiol. Biochem. 2018;49(3):1197–1207. doi: 10.1159/000493301. [DOI] [PubMed] [Google Scholar]
  • 99.Macri V., Mahida S.N., Zhang M.L., et al. A novel trafficking-defective HCN4 mutation is associated with early-onset atrial fibrillation. Heart Rhythm. 2014;11(6):1055–1062. doi: 10.1016/j.hrthm.2014.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 100.Biel S., Aquila M., Hertel B., et al. Mutation in S6 domain of HCN4 channel in patient with suspected Brugada syndrome modifies channel function. Pflugers Arch. 2016;468(10):1663–1671. doi: 10.1007/s00424-016-1870-1. [DOI] [PubMed] [Google Scholar]
  • 101.Baruscotti M., Bianco E., Bucchi A., DiFrancesco D. Current understanding of the pathophysiological mechanisms respon-sible for inappropriate sinus tachycardia: role of the If “fun-ny” current. J. Interv. Card. Electrophysiol. 2016;46(1):19–28. doi: 10.1007/s10840-015-0097-y. [DOI] [PubMed] [Google Scholar]
  • 102.Baruscotti M., Bucchi A., Milanesi R., et al. A gain-of-function mutation in the cardiac pacemaker HCN4 channel increasing cAMP sensitivity is associated with familial Inappropriate Si-nus Tachycardia. Eur. Heart J. 2017;38(4):280–288. doi: 10.1093/eurheartj/ehv582. [DOI] [PubMed] [Google Scholar]
  • 103.Schweizer P.A., Schröter J., Greiner S., et al. The symptom complex of familial sinus node dysfunction and myocardial noncompaction is associated with mutations in the HCN4 channel. J. Am. Coll. Cardiol. 2014;64(8):757–767. doi: 10.1016/j.jacc.2014.06.1155. [DOI] [PubMed] [Google Scholar]
  • 104.Weigl I., Geschwill P., Reiss M., et al. The C-terminal HCN4 variant P883R alters channel properties and acts as genetic modifier of atrial fibrillation and structural heart disease. Biochem. Biophys. Res. Commun. 2019;519(1):141–147. doi: 10.1016/j.bbrc.2019.08.150. [DOI] [PubMed] [Google Scholar]
  • 105.Nawathe P.A., Kryukova Y., Oren R.V., et al. An LQTS6 MiRP1 mutation suppresses pacemaker current and is associated with sinus bradycardia. J. Cardiovasc. Electrophysiol. 2013;24(9):1021–1027. doi: 10.1111/jce.12163. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 106.Bucchi A., Baruscotti M., Nardini M., et al. Identification of the molecular site of ivabradine binding to HCN4 channels. PLoS One. 2013;8(1):e53132. doi: 10.1371/journal.pone.0053132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 107.Bucchi A., Tognati A., Milanesi R., Baruscotti M., DiFrancesco D. Properties of ivabradine-induced block of HCN1 and HCN4 pacemaker channels. J. Physiol. 2006;572(Pt 2):335–346. doi: 10.1113/jphysiol.2005.100776. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 108.Novella Romanelli M., Sartiani L., Masi A., et al. HCN channels modulators: the need for selectivity. Curr. Top. Med. Chem. 2016;16(16):1764–1791. doi: 10.2174/1568026616999160315130832. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 109.Delpón E., Valenzuela C., Pérez O., et al. Mechanisms of block of a human cloned potassium channel by the enantiomers of a new bradycardic agent: S-16257-2 and S-16260-2. Br. J. Pharmacol. 1996;117(6):1293–1301. doi: 10.1111/j.1476-5381.1996.tb16728.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 110.Haechl N., Ebner J., Hilber K., Todt H., Koenig X. Pharmacolog-ical profile of the bradycardic agent ivabradine on human cardiac ion channels. Cell. Physiol. Biochem. 2019;53(1):36–48. doi: 10.33594/000000119. [DOI] [PubMed] [Google Scholar]
  • 111.Mengesha H.G., Tafesse T.B., Bule M.H. If channel as an emerg-ing therapeutic target for cardiovascular diseases: a review of current evidence and controversies. Front. Pharmacol. 2017;8:874. doi: 10.3389/fphar.2017.00874. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 112.Fainzilber M, Nakamura T, Lodder JC, Zlotkin E, Kits KS, Burlingame AL. γ-Conotoxin-PnVIIA, a γ-carboxyglutamate-containing peptide agonist of neuronal pacemaker cation currents. Biochemistry. 1998;37(6):1470–1477. doi: 10.1021/bi971571f. [DOI] [PubMed] [Google Scholar]
  • 113.Agarwal A., Kumar T., Ravindranath K.S., Bhat P., Manjunath C.N., Agarwal N. Sinus node dysfunction complicating viper bite. Asian Cardiovasc. Thorac. Ann. 2015;23(2):212–214. doi: 10.1177/0218492313501819. [DOI] [PubMed] [Google Scholar]
  • 114.Meng Q.T., Xia Z.Y., Liu J., Bayliss D.A., Chen X. Local anes-thetic inhibits hyperpolarization-activated cationic currents. Mol. Pharmacol. 2011;79(5):866–873. doi: 10.1124/mol.110.070227. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 115.Putrenko I., Yip R., Schwarz S.K.W., Accili E.A. Cation and volt-age dependence of lidocaine inhibition of the hyperpolariza-tion-activated cyclic nucleotide-gated HCN1 channel. Sci. Rep. 2017;7(1):1281. doi: 10.1038/s41598-017-01253-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 116.Zhang X., Zhao Y., Du Y., et al. Effect of ketamine on mood dysfunction and spatial cognition deficits in PTSD mouse models via HCN1-BDNF signaling. J. Affect. Disord. 2021;286:248–258. doi: 10.1016/j.jad.2021.02.058. [DOI] [PubMed] [Google Scholar]
  • 117.McClure K.J., Maher M., Wu N., et al. Discovery of a novel series of selective HCN1 blockers. Bioorg. Med. Chem. Lett. 2011;21(18):5197–5201. doi: 10.1016/j.bmcl.2011.07.051. [DOI] [PubMed] [Google Scholar]
  • 118.Cao Y., Chen S., Liang Y., et al. Inhibition of hyperpolariza-tion-activated cyclic nucleotide-gated channels by -blocker carvedilol. Br. J. Pharmacol. 2018;175(20):3963–3975. doi: 10.1111/bph.14469. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 119.Chen S-J., Xu Y., Liang Y-M., et al. Identification and charac-terization of a series of novel HCN channel inhibitors. Acta Pharmacol. Sin. 2019;40(6):746–754. doi: 10.1038/s41401-018-0162-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 120.DiFrancesco D. Some properties of the UL-FS 49 block of the hyperpolarization-activated current (i(f)) in sino-atrial node myocytes. Pflugers Arch. 1994;427(1-2):64–70. doi: 10.1007/BF00585943. [DOI] [PubMed] [Google Scholar]
  • 121.Del Lungo M., Melchiorre M., Guandalini L., et al. Novel blockers of hyperpolarization-activated current with isoform selectivity in recombinant cells and native tissue. Br. J. Pharmacol. 2012;166(2):602–616. doi: 10.1111/j.1476-5381.2011.01782.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 122.Melchiorre M., Del Lungo M., Guandalini L., et al. Design, synthesis, and preliminary biological evaluation of new iso-form-selective f-current blockers. J. Med. Chem. 2010;53(18):6773–6777. doi: 10.1021/jm1006758. [DOI] [PubMed] [Google Scholar]
  • 123.Romanelli M.N., Del Lungo M., Guandalini L., et al. EC18 as a tool to understand the role of HCN4 channels in mediating hyperpolarization-activated current in tissues. ACS Med. Chem. Lett. 2019;10(4):584–589. doi: 10.1021/acsmedchemlett.8b00587. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 124.Dini L., Del Lungo M., Resta F., et al. Selective blockade of HCN1/HCN2 channels as a potential pharmacological strategy against pain. Front. Pharmacol. 2018;9:1252. doi: 10.3389/fphar.2018.01252. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 125.Du Y, Zhang J, Xi Y, et al. β1-Adrenergic blocker bisoprolol reverses down-regulated ion channels in sinoatrial node of heart failure rats. J. Physiol. Biochem. 2016;72(2):293–302. doi: 10.1007/s13105-016-0481-9. [DOI] [PubMed] [Google Scholar]
  • 126.Okamoto T., Harnett M.T., Morikawa H. Hyperpolarization-activated cation current (Ih) is an ethanol target in midbrain dopamine neurons of mice. J. Neurophysiol. 2006;95(2):619–626. doi: 10.1152/jn.00682.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 127.Chen Y., Wu P., Fan X., et al. Ethanol enhances human hy-perpolarization-activated cyclic nucleotide-gated currents. Alcohol. Clin. Exp. Res. 2012;36(12):2036–2046. doi: 10.1111/j.1530-0277.2012.01826.x. [DOI] [PubMed] [Google Scholar]
  • 128.Nakashima K., Nakao K., Matsui H. Discovery of novel HCN4 blockers with unique blocking kinetics and binding proper-ties. SLAS Discov. 2021;26(7):896–908. doi: 10.1177/24725552211013824. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 129.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(12):E4260. doi: 10.3390/ijms21124260. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 130.Chen H., Chen Y., Yang J., Wu P., Wang X., Huang C. Effect of Ginkgo biloba extract on pacemaker channels encoded by HCN gene. Herz. 2020 doi: 10.1007/s00059-020-04933-z. [DOI] [PubMed] [Google Scholar]
  • 131.Satoh H. Suppression of pacemaker activity by Ginkgo biloba extract and its main constituent, bilobalide in rat sino-atrial nodal cells. Life Sci. 2005;78(1):67–73. doi: 10.1016/j.lfs.2005.04.081. [DOI] [PubMed] [Google Scholar]
  • 132.Satoh H., Nishida S. Electropharmacological actions of Ginkgo biloba extract on vascular smooth and heart muscles. Clin. Chim. Acta. 2004;342(1-2):13–22. doi: 10.1016/j.cccn.2003.12.014. [DOI] [PubMed] [Google Scholar]
  • 133.DiFrancesco D. A brief history of pacemaking. Front. Physiol. 2020;10:1599. doi: 10.3389/fphys.2019.01599. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 134.Shin K.S., Rothberg B.S., Yellen G. Blocker state dependence and trapping in hyperpolarization-activated cation channels: evidence for an intracellular activation gate. J. Gen. Physiol. 2001;117(2):91–101. doi: 10.1085/jgp.117.2.91. [DOI] [PMC free article] [PubMed] [Google Scholar]

Articles from Current Cardiology Reviews are provided here courtesy of Bentham Science Publishers

RESOURCES