Next Article in Journal
M3 Receptor Pathway Stimulates Rapid Transcription of the CB1 Receptor Activation through Calcium Signalling and the CNR1 Gene Promoter
Next Article in Special Issue
Piezo1 Activation Prevents Spheroid Formation by Malignant Melanoma SK-MEL-2 Cells
Previous Article in Journal
Absence of Scaffold Protein Tks4 Disrupts Several Signaling Pathways in Colon Cancer Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Flavonoids as Modulators of Potassium Channels

by
Monika Richter-Laskowska
1,
Paulina Trybek
2,
Domenico Vittorio Delfino
3 and
Agata Wawrzkiewicz-Jałowiecka
4,*
1
The Centre for Biomedical Engineering, Łukasiewicz Research Network—Krakow Institute of Technology, 30-418 Krakow, Poland
2
Faculty of Science and Technology, University of Silesia in Katowice, 41-500 Chorzów, Poland
3
Department of Internal Medicine, Università degli Studi di Perugia, 06123 Perugia, Italy
4
Department of Physical Chemistry and Technology of Polymers, Silesian University of Technology, 44-100 Gliwice, Poland
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(2), 1311; https://doi.org/10.3390/ijms24021311
Submission received: 9 November 2022 / Revised: 2 January 2023 / Accepted: 4 January 2023 / Published: 9 January 2023
(This article belongs to the Special Issue Modulation of Ion Channels)

Abstract

:
Potassium channels are widely distributed integral proteins responsible for the effective and selective transport of K + ions through the biological membranes. According to the existing structural and mechanistic differences, they are divided into several groups. All of them are considered important molecular drug targets due to their physiological roles, including the regulation of membrane potential or cell signaling. One of the recent trends in molecular pharmacology is the evaluation of the therapeutic potential of natural compounds and their derivatives, which can exhibit high specificity and effectiveness. Among the pharmaceuticals of plant origin, which are potassium channel modulators, flavonoids appear as a powerful group of biologically active substances. It is caused by their well-documented anti-oxidative, anti-inflammatory, anti-mutagenic, anti-carcinogenic, and antidiabetic effects on human health. Here, we focus on presenting the current state of knowledge about the possibilities of modulation of particular types of potassium channels by different flavonoids. Additionally, the biological meaning of the flavonoid-mediated changes in the activity of K + channels will be outlined. Finally, novel promising directions for further research in this area will be proposed.

1. Introduction

The potassium channels are transmembrane tetrameric proteins allowing for a fast (over 10 6 ions per second) and selective transport of K + ions across the biological membranes down their electrochemical gradient [1]. Although the K + channels are related members of one protein family, they can be divided into several groups according to the structural and mechanistic differences [2,3,4,5]. First, one can discern the voltage-regulated channels (Kv) and the C a 2 + -regulated K + channels (KCa), being structurally distinguished by six transmembrane domains (TMDs). Another group is formed by the “leak”, which is the double pore K + channels (K2P) composed of four TMDs. The last group is constituted by the inward rectifier potassium channels (Kir), having two TMDs.
The activity of the K + channels plays a pivotal role in a myriad of cellular processes, for instance, volume regulation, proliferation, hormone secretion, neurotransmitter release, and modulation of potential in electrically excitable and non-excitable cells [1]. Due to their fundamental physiological roles, they appear promising drug targets for many diseases ranging from cardiovascular, metabolic, and neurodegenerative disorders to cancers [6,7,8,9,10,11]. The potential therapeutic benefits of the potassium channels targeting by pharmacological agents depend on the efficacy, strength, and selectivity of the interaction between an active compound and a particular ion channel subtype. It is also challenging to introduce such a channel modulator, which would be delivered or act only within a pathological tissue and would not exhibit any harmful side effects.
The mentioned issues give rise to a growing trend in the rational design of novel K + channels’ regulators which requires multidisciplinary investigations among different scientific areas, such as medicinal chemistry, molecular biochemistry, bioinformatics, big data processing, bioengineering, genomics, proteomics, and metabolomics [7,11]. A promising approach is to improve the pharmacokinetics of existing channel-modulating substances such as antibodies, venom peptides, nutraceuticals or medicinal plants [11,12] belonging to the traditional medicine defined by the WHO as “...the sum total of knowledge, skills and practices based on the theories, beliefs and experiences indigenous to different cultures that are used to maintain health as well as to prevent, diagnose, improve or treat physical and mental illnesses...” [13].
Throughout human history, plants and natural products have been the source of effective agents to cure illnesses and improve health. These plant-based traditional curative systems are still widely used and continue to play an essential role in primary health care. Traditional medicines over the years have proved to be an invaluable guide in the current screening of bioactive molecules for therapeutic applications [14] and still are widely used as modern pharmaceuticals or health-beneficial nutrients [15,16,17,18,19]. From a chemical point of view, natural products can be classified based on their biosynthetic origin as (1) polyketides, (2) shikimic-acid-derived natural products, (3) terpenes, (4) glycosides and (5) alkaloids. Among them, we have focused our attention on flavonoids belonging to the group of shikimic-acid-derived natural products [20].
Flavonoids comprise a wide group of polyphenolic compounds of plant origin, which exhibit anti-oxidative, anti-inflammatory, anti-mutagenic, and anti-carcinogenic properties, as well as the capability to regulate the functioning of key cellular enzymes [21,22]. Flavonoids can be divided into subclasses such as flavones, flavonols, flavanols, flavanones, isoflavones, anthocyanidins, and chalcones. Due to their ability to modulate cell physiology, they have a clinically proven positive impact in counteracting a plethora of health problems such as cardiovascular and metabolic diseases, or other inflammation-related pathologies, including cancers, as summarized in [17].
In this work, we will enlighten the molecular aspects of the impact of flavonoids on living cells via the K + channels’ modulation. The interactions of flavonoid molecules with specific subtypes of potassium channels (direct or indirect), and the consequent alterations of their transport properties, can be a crucial factor underlying the observed changes in flavonoid-stimulated cells’ biology. It inspired us to summarize in this review the current state of knowledge about the impact of flavonoids on the activity of potassium channels. This work is arranged according to the main subgroups of these channels. It can provide a valuable contribution from the perspective of indication of the most promising channel-targeting flavonoids and analysis of their derivatives in future research. This overview also outlines the promising directions for further in-depth analysis, considering the physiological effects of flavonoid-induced changes in K + channels’ activity.

2. Kv Channels

Voltage-gated potassium (Kv) channels are the largest ion channel family in the human organism consisting of 40 members divided into 12 subfamilies Kv1–Kv12. They have many physiological functions such as shaping action potentials, maintenance of membrane potential, neuronal repolarization, modulation of C a 2 + signaling, cell volume regulation and control of cellular proliferation and migration [23,24]. Depending on their functionality and location, they are able to pass the current in and out of a cell in response to a change in transmembrane electric potential.
In the following subsections, we discuss the influence of flavonoids on the activity of particular subtypes of the Kv channels. The key information is presented in Table 1 and Table 2. Table 2 refers only to the modulation of the Kv 11.1 channels. Thus, it is presented in the Section 2.5, which describes the effects of flavonoids on these channels.

2.1. Kv1.3 Channel

Kv1.3 channels are activated, as all the other voltage-gated potassium channels, by a change of the membrane potential (membrane depolarization). Although they are expressed in many different tissues of the human body, they play the most prominent role in the T lymphocytes [50,51] where they are responsible for cell activation. Accordingly, the inhibition of these ion channels results in the inactivation of the T cells, which promotes immune suppression. It makes the Kv1.3 channel a therapeutic target for the treatment of such diseases as sclerosis, type 1 diabetes, and rheumatoid arthritis [6,50,52]. The recent findings suggest that an increased expression of these channels is required to induce apoptosis in the cancer cells. Therefore, the Kv1.3 channels are a new molecular target in both the diagnostics and therapy of some oncological diseases (e.g., breast cancer, colon cancer) [53].
Throughout the years, the group of Teisseyre investigated the impact of a large number of different flavonoids on activation of the Kv1.3 channels, both in normal and cancer T cells [53]. They discovered the inhibitory effects of genistein, a popular tyrosine kinase inhibitor, which suppresses the activation of the Kv1.3 channel in a concentration-dependent manner with the half-blocking range of I C 50 = 30–60 μ M [25]. These results are consistent with the previous study presented in [54]. Moreover, Teisseyre et al. observed that under the influence of daidzein, which is a structural analog of genistein, the ion channel’s activity remains unaffected. They further demonstrated that resveratrol (a non-flavonoid polyphenol) is able to decrease the activity of the Kv1.3. channel. This effect is slowly reversible, and it is exerted in a concentration-dependent manner. At the same time, it turned out that the co-application of this polyphenol with genistein did not significantly change the suppression effect of resveratrol. Since in the previous study [25], it was demonstrated that genistein has a similar impact on the Kv1.3 channel as resveratrol, the authors suggested that the inhibitory effects of these compounds are independent of each other, and they interact with different binding sites of the channel.
In the following experiments [30], the influence of four flavonoids: aromadendrin, naringenin and its two derivatives, naringenin-4 ,7-dimethylether and naringenin-7-methyl-ether, on the Kv1.3 channel activity was investigated in human T lymphocytes isolated from peripheral blood. The single-channel patch-clamp traces revealed that naringenin and aromadendrin did not reduce the Kv1.3 current at the concentration of c = 30 μ M. On the contrary, the two investigated derivatives of naringenin enabled us to reduce the ionic currents at the same concentrations, and the most effective inhibitory effect was observed for the naringenin-4 ,7-dimethylether. The authors suggested that the suppressing capabilities of the naringenin-related methylated compounds are due to the presence of one or two methoxyl groups in their structure, which possibly interact in some, yet unknown, way with the Kv1.3 channel protein. Nevertheless, although the methylated versions of naringenin—naringenin-4 ,7-dimethylether and naringenin-7-methylether—are quite effective blockers of the Kv1.3 channel, they do not enable the complete channel inhibition [30].
In contrast, in [32], it was discovered that 8-prenylnaringenin blocks completely the Kv1.3 channel at c = 10 μ M. The presence of the prenyl group is anticipated to promote the inhibitory abilities of this compound. Indeed, a few years later, it was demonstrated that other compounds of such structural characteristics, xanthohumol and isoxanthohumol, are able to effectively suppress the activity of the Kv1.3 channel [31]. Although xanthohumol turned out to be slightly more effective in decreasing the ion current, the administration of neither of the investigated compounds led to a complete channel inhibition at the concentration c = 30 μ M. Even though xanthohumol and isoxanthohumol are less potent in inhibition of the Kv1.3 channel than 8-prenylnaringenin, they are much more effective than the other natural plant-derived compounds such as genistein and resveratrol. Therefore, these results confirm the hypothesis that prenylated flavonoids are much more effective in blocking the Kv1.3 channel due to the presence of the prenyl group, which facilitates the non-conducting state of the channel. To further investigate this problem, Teisseyere and his collaborators decided to make a comparative study between natural, plant-derived flavonoids and those which possess also a prenyl group in their structure [26]. They found that pure flavonoids such as baicalein, wogonin, and luteolin were ineffective. In contrast, two other non-prenylated compounds, acacetin and chrysin, were able to suppress the single-channel current regardless of the absence of the prenyl group in its structure. However, the most potent inhibition was obtained during the application of the 6-prenylnaringenin (6–PR), which suppresses the ion channel in a concentration-dependent manner with I C 50 = 5.8 μ M . Once again, it suggested that the mechanism responsible for the effective inactivation of the channel is based on interactions of the prenyl group with the channel protein. The authors compared the obtained results regarding the modulation of the Kv1.3 by acacetin with another similar work published by Zhao et al. [27]. They found that there exists a significant discrepancy between their findings and those obtained in [27]. According to Teisseyere and his collaborators, the half-blocking concentration for this compound is approximately equal to I C 50 = 30 μ M . In contrast, in [27], it was reported that I C 50 = 21 μ M . The authors of [26] attributed this difference to the time of the ion channel exposition to this flavonoid. In the experiment performed by the Teisseyere group, the incubation in acacetin dwelled for no longer than 5 min, and its effect on the channel was fully reversible. On the contrary, Zhao and his collaborators exposed the Kv1.3 channel to this compound for at least 15 min, and its effects turned out to be only partially reversible. It suggests that the inhibitory effect of acacetin can change with time, and it may exhibit cytotoxicity at higher concentrations. The wide selection of the examined compounds in the work of Teisseyre [26] allowed the authors to draw the conclusion that there is no correlation between the inhibitory abilities of the flavonoids and their cytotoxicity (at least at the investigated type of cells).
In the most recent study [28], the Teisseyere group investigated the impact of the different flavonoids co-applied with statins, which alone had turned out to be effective in blocking the Kv1.3 channel [55]. Their findings showed that in most cases, modulation of the ion channel with a flavonoid accompanied by a statin is more potent than the administration of this flavonoid alone. The results also demonstrated that the inhibitory effects are not always additive. Therefore, the mechanism of the ion channel’s modulation is complex, and the observed channel’s transport capability strongly depends on the chosen proportions in a mixture of statins and flavonoids. The blocking effects of simvastatin and mevastatin co-applied with 8-prenylnaringenin and simvastatin with 6-prenylnaringenin were significantly more potent than predicted by the simple additive model. On the contrary, the inhibition of simvastatin with xanthohumol and acacetin turned out to be notably weaker than expected by a simple addition.
Another group of flavonoids, chalcones, being derivatives of khellinone comprising two aryl rings linked by an α , β -unsaturated ketone, also exert the inhibitory effect on the Kv1.3 channel. Although the khellinone itself cannot be considered a very potent inhibitor of this ion channel with the half-blocking concentration of I C 50 = 45 μ M , the I C 50 is substantially lower for its dimers [56,57]. Thus, for clarity, khellinone and its dimer are not elucidated in Table 1. Nevertheless, the complete summary of the Kv1.3 inhibiting properties of this compound and its derivatives is clearly presented in Table 1 in [57]. Considering other chalcones, recent studies show that also Licochalcone A is effective in blocking the Kv1.3 channel [33]. Only the concentration of c 0.8 μ M was enough to reduce the ion channel activity by half.

2.2. Kv1.5 Channel

The channels belonging to the Kv1.5 family are expressed in many tissues of the human body [58,59]. The greatest attention of the scientists is, however, focused on its expression in the heart. It has been demonstrated [60,61] that the Kv1.5 channel conducts the ultra-rapid delayed rectifier current I K u r , which plays an important role in shaping the atrial action potential (AP) repolarization [62]. The studies suggest that the inhibition of this channel can contribute to the prolongation of the AP duration and, by this, stop the atrial fibrillation (AF) [63]. It is important to note that although the Kv1.5 channel is present in the atria, it is not expressed in the ventricular muscle in the heart. These two factors, i.e., provision of the current driving the AP and selective expression of this ion channel (its presence in the atria and absence in the ventricle), consider the Kv1.5 channels as a potential target for the treatment of the cardiac arrhythmia [64,65,66]. For this reason, it would be useful to find new efficient inhibitors of the Kv1.5 channel.
In [35], Wang et al. studied the inhibitory effects of the hesperetin on the I K u r through the Kv1.5 channels expressed in the HEK 293 cells. It turned out that although externally applied hesperetin can significantly suppress the ultra-rapid delayed rectifier K + current in the concentration-dependent manner with I C 50 = 23.15 μ M, the presence of this flavonoid in pipette solution yielded no effect. The authors concluded that hesperetin interacts with a channel protein only from the exterior. They also observed inhibition of the I K u r along with the suppression of the Kv1.5 channel. However, the study does not provide any information about the action of this flavonoid on the atrial action potential. Thus, it remains unclear whether hesperetin can induce the termination of atrial fibrillation. Another study [38] investigated the impact of hesperetin on the expression of the Kv1.5 channels in coronary arteries of diabetic and non-diabetic rats. The known fact is that diabetes downregulates the expression of these channels in arterial myocytes. Therefore, it would be beneficial to find a biochemical agent which enhances its expression to the proper level. According to [38], although hesperetin has no impact on the expression of Kv1.5 ion channels in arterial myocytes, it increases the expression of Kv1.2 channels which is also desirable during the treatment of diabetic patients.
In [36], Yang et al. investigated the effect of quercetin on I K u r conducted by the wild-type (WT) and mutant (1502A) human Kv1.5 channels. The measured traces obtained from the patch–clamp experiment revealed no effect of quercetin on the mutated Kv1.5 channels’ functioning. However, significant enhancement of ionic currents in the presence of quercetin was observed for the WT channels (with E C 50 = 37   μ M). These findings allowed us to conclude that quercetin binds preferentially to neutral amino acid I502, which is located in the S6 helix of the Kv1.5 channel. The observed increase in activation of the Kv1.5 cannot be beneficial in the AF treatment during the early phase of this disease. However, it turns out that the chronic AF results in the reduced expression of Kv1.5 α subunits and prolongation of the action potential [67]. It is, therefore, crucial to increase the I K u r at this stage of the disease. From this perspective, quercetin may be used in the treatment of late phase and chronic AF due to its ability to activate the Kv1.5 channels. Another study [37] showed that quercetin can reverse the inhibition of the Kv1.5 current, primarily induced by the monocrotaline, which generates pulmonary arterial hypertension (PAH) in rats. It indicates another potential application of this flavonoid in treatment of patients suffering from cardiovascular diseases. In turn, very weak inhibitory effects of quercetin were demonstrated in [38] based on the analysis of the Kv1.5 channel stimulation by quercetin and its methylated derivatives (3,7,3 ,4 -tetramethylquecertin, 3,5,7,3 ,4 -pentamethylquecertin) in HEK 293 cells. It turns out that suppression of the Kv1.5 channel activity by 3 and 10 μ M quercetin is only weak by ≈ 3.0 ± 1.8 % and ≈ 5.2 ± 3.1 % , respectively. Effects of 3,5,7,3 ,4 -pentamethylquecertin are similar to those induced by pure quercetin; i.e., at 3 and 10 μ M, it decreased the current by 3.4 ± 2.4% and 8.3 ± 2.5%. The situation notably changes in the case of the application of 3,7,3 ,4 -tetramethylquecertin at 3 and 10 μ M: it inhibited the current by 12.1 ± 2.2% and 20.5 ± 5.2%.
In [39], Choi and his colleagues decided to analyze the impact of (−)-epigallocatechin-3-gallate (EGCG) (the main polyphenolic component of green tea) on the ion currents through Kv1.5 channels expressed in Chinese hamster ovary cells (CHO). They observed the significant downregulation of this ion channel in the presence of this flavonoid, which occurs in a concentration-dependent manner. As it turned out, this inhibition was not suppressed by the protein tyrosine kinase, tyrosine phosphatase and protein kinase C inhibitors. The more profound kinetic analysis of the currents revealed that EGCG interacts directly with multiple states (conformations) of the Kv1.5 channel. It preferentially binds to the channel in the closed state, and blocks it by pore occlusion during the depolarization.
Noguchi et al. [40] examined the inhibitory effect of isoliquiritigenin (ISL) flavonoid contained in licorice on the Kv1.5 channel expressed in the CHO cell line. They observed the mediatory suppression effect of this compound on this ion channel. The application of 100 μ M ISL at a membrane potential of 40 mV inhibited the Kv1.5 I K u r current by 38.3%.
In [68], the authors investigated the effect of apigenin on pulmonary hypertension. Although this natural compound does not affect the value of a Kv1.5 ion current, it increases its expression in the pulmonary artery smooth muscle cells (PASMC) of hypoxia-exposed rats, stimulating their apoptosis. These results provide promising therapeutic targets for the treatment of pulmonary hypertension. Another study analyzing results obtained from the patch-clamp experiment performed on HEK 293 cells revealed the weak inhibitory effect of the apigenin on the Kv1.5 channel in the presence of c = 3   μ M ( 4.6 ± 2.5 % ) and c = 10   μ M ( 11.1 ± 2.9 % ) of this flavonoid. Larger effects were observed for the double methylated compound, 7,4 -dimethylapigenin, which can suppress the ion current by 16.4 ± 3.1% and 28.8 ± 6.0% at the same concentrations. However, the major reduction of the channel’s activity was observed after application of 5,7,4 -trimethylapigenin, which occured in a concentration-dependent manner with I C 50 = 6.2 μ M . In that case, the current was inhibited by 28.9 ± 2.4% with c = 3   μ M and 70.2 ± 2.8% with c = 10   μ M. The more profound analysis of the blocking properties of this flavonoid revealed that it binds mainly to open channels. The inhibition efficacy of 5,7,4 -trimethylapigenin on hKv1.5 was also confirmed in the human atrial myocytes I K u r , which suggests that the Kv1.5 α subunit is the dominant target for the drug (channel blocker) needed in the treatment of atrial fibrillation.
In [42], Li et al. noted the inhibitory impact of the acacetin on the ultra-rapid delayed rectifier current I K u r and Kv1.5 current during the patch-clamp experiment performed on the atrial myocytes. They concluded that acacetin decreases the I K u r and downregulates other important cardiac currents (such as transient outward I t o and acetylcholine-activated I K A C h K + currents), which altogether have a significant impact on the prolongation of the action potential. Several years later, they carried out the analogous experiment on the human HEK 293 cell line [41], which is best suited for investigations on the molecular mechanisms of channel binding with flavonoids. They once again confirmed the inhibitory properties of acacetin, which blocks the I K u r in a use- and frequency-dependent manner. They found that this flavonoid binds to channels in their open or closed conformations. The acacetin-mediated blocking of the open hKv1.5 channels is mediated by binding this flavonoid to the S6 channel domain.
In [34], it is reported that myricetin can exert beneficial anti-arrhythmic effects via Kv1.5 channel regulation. The experiments on the HEK 293 cells showed that this drug enables effective blocking of the channel and inhibits the I K u r .

2.3. Kv2.1 Channel

The voltage-dependent potassium channels Kv2.1 are expressed both in the central and peripheral nervous system of mammals where they are predominant mediators of the delayed rectifier current [69,70,71]. They play a prominent role in shaping neuronal excitability [72] and in the glucose-stimulated insulin secretion [73], which makes this channel a promising target in treatment of diabetes.
In [46], the authors showed an inhibitory impact of genistein on the activation of the Kv2.1 channels. The analysis of the data obtained from the patch-clamp technique on HEK293 cells revealed that in the presence of this compound, the ion channel is inhibited in a concentration-dependent manner. The more profound kinetic analysis showed that genistein shifted the voltage dependence of channels’ activation and inactivation to membrane hyperpolarization. It also accelerated the closed-state inactivation and delayed the recovery from inactivation.
In [47], Gu with his colleagues demonstrated the inhibitory impact of (−)-naringenin 4,7-dimethyl ether ((−)-NRG-DM) on the Kv2.1 channel expressed in the CHO cells. They reported that it suppresses the ion current in a concentration-dependent manner with I C 50 21 μ M and shifts half-maximal voltage toward the higher potentials.
The authors of the article [37] found that, similarly as in the case of Kv1.5 channels, quercetin can prevent the inhibition of the Kv2.1 currents in the pulmonary artery smooth muscle cells (PASMC) of rats pretreated with the monocrotaline.
In [40], the authors studied the impact of licorice and isoliquiritigenin on I K u r mediated by Kv2.1 expressed in H9c2 cells derived from rat cardiac myoblasts. They discovered the blocking impact of this compound on the I K u r with I C 50 = 0.11 μ M , making it one of the strongest inhibitors of this current.

2.4. Kv4 Channels

The Kv4 (Shal) channels are widely expressed in the neurons of different animals [74,75,76]. They mediate the fast A-type K + currents and are thought to be responsible for the fundamental electrical properties of nerve cells.
There is not much information about the modulation of these types of ion channels by the flavonoids. In [77,78], it was shown that pinocembrin upregulates the expression of Kv4.2 channels, which can be beneficial in the treatment of ventricular arrhythmia. A little bit more is known about Kv4.3, which according to [48] can be blocked by genistein and, to a smaller extent, by daidzein. The analysis of the Kv4.3 recordings obtained by the patch-clamp technique in the CHO cells revealed that genistein inhibits the current in a reversible and concentration-dependent manner with I C 50 125 μ M . Moreover, it was found that this inhibition is direct: genistein downregulates the activity of the Kv4.3 channel by binding to the closed-inactivated state of the channel, and these interactions are definitely not mediated by the protein tyrosine kinase mechanism. This research group also investigated the impact of the other flavonoids on the Kv4.3 channel: daidzein and genistin, which are the structural analogs of genistein. Their results showed that although daidzein was able to downregulate the channel in a concentration-dependent manner, the complete inhibition could not be achieved. On the other hand, the presence of genistin had no effect on the activation of this ion channel. Another study [38] showed that 5,7,4 -trimethylapigenin is an effective blocker of the Kv4.3 channels in the human atrial myocytes. Thus, it may contribute to the prolongation of the atrial action potential duration needed for the treatment of the atrial fibrillation. The other studies revealed that also (−)-epigallocatechin-3-gallate [44] and naringenin [45] exert a mild inhibitory impact on the Kv4.3 channels.

2.5. hERG Channels

The ether-à-go-go-related hERG ion channel (Kv11.1 channel), similarly to the already mentioned Kv1.5 channel, is responsible for the electrical activity of the heart. It regulates the cardiac action potential by mediating the repolarizing current [79,80].
The impact of different flavonoids on the hERG channels (Kv11.1 channels) has been already discussed in the review [81]. Here, we will briefly summarize the information gathered by the authors of that paper and supplement it with the recent advancements in this field (Table 2).
One of the most important studies concerning the modulation of the hERG channel was performed by Zitron et al. [82] in 2005, who screened a large number of flavonoids for their inhibitory abilities. Based on the patch-clamp recordings, they found that the most potent inhibitors of the Kv11.1 channels are naringenin, morin, and hesperetin. Naringenin blocked the hERG channel expressed both in the Xenopus laevis oocytes and HEK 293 cells with half-blocking concentrations I C 50 = 102 μ M and I C 50 = 36.5 μ M , respectively. A more detailed analysis revealed that the channels are blocked in the open and inactivated states by naringenin but not in the closed states [83]. Moreover, the ECG examinations suggested that the blockade of the hERG channel induced by naringenin results in prolongation of the QT interval [45]. Other authors showed that the addition of the antiarrhythmic drugs can strenghten the influence of naringenin on the hERG channel [84]. Nevertheless, such a combination may have an overstimulating effect and pose increased risk of arrhythmias.
Table 2. The effects of different flavonoids on the activity of Kv11.1 (hERG) channels. E C 50 is the concentration of a flavonoid that gives a half-maximal response. I C 50 is the concentration of a flavonoid concentration at 50% channel inhibition. The arrows symbolize the type of observed effects on the channel activity: ↓ inhibition, ↑ activation, → no effect. The table is in most part adapted from [81] with permission from Elsevier (2023).
Table 2. The effects of different flavonoids on the activity of Kv11.1 (hERG) channels. E C 50 is the concentration of a flavonoid that gives a half-maximal response. I C 50 is the concentration of a flavonoid concentration at 50% channel inhibition. The arrows symbolize the type of observed effects on the channel activity: ↓ inhibition, ↑ activation, → no effect. The table is in most part adapted from [81] with permission from Elsevier (2023).
FlavonoidType of CellEffect I C 50 / E C 50 References
AcacetinHEK 293 32.4 μ M Li et al. [42] (2008)
ApigeninXenopus oocyte Zitron et al. [82] (2005)
ChrysinXenopus oocyte Zitron et al. [82] (2005)
DaidzeinHEK 293 Zhang et al. [85] (2008)
7,8-DimethoxyflavoneXenopus oocyte Du et al. [86] (2015)
(−)-Epigallocatechin gallateHEK 293 6 μ M Kelemen et al. [87] (2007)
Xenopus oocyte 20.5 μ M
CHO Kang et al. [44] (2010)
FisetinXenopus oocyte Zitron et al. [82] (2005)
HEK 293 38.4 μ M Sun et al. [88] (2017)
FlavoneXenopus oocyte Zitron et al. [82] (2005)
GalanginXenopus oocyte Zitron et al. [82] (2005)
HEK 293 Sun et al. [88] (2017)
GenisteinHEK 293 Zhang et al. [85] (2008)
HesperetinXenopus oocyte 289 μ M Zitron et al. [82] (2005)
267 μ M Scholz et al. [89] (2007)
HesperidinXenopus oocyte Zitron et al. [82] (2005)
IsorhamnetinHEK 293 Sun et al. [88] (2017)
KaempferolXenopus oocyte Zitron et al. [82] (2005)
HEK 293 Sun et al. [90] (2017)
LiquiritigeninCHL 53 μ M Sweeney et al. [91] (2019)
LuteolinHEK 293 Sun et al. [88] (2017)
7,8-MethylenedioxyflavoneXenopus oocyte Du et al. [86] (2015)
MorinXenopus oocyte 111 μ M Zitron et al. [82] (2005)
HEK 293 Sun et al. [88] (2017)
MyricetinXenopus oocyte Zitron et al. [82] (2005)
HEK 293 Sun et al. [88] (2017)
NaringeninXenopus oocyte 103 μ M Scholz et al. [83] (2005)
173 μ M Lin et al. [84] (2008)
102 μ M Zitron et al. [82] (2005)
HEK 293 36.5 μ M
CHO 35 μ M Sanson et al.[45] (2022)
NaringinXenopus oocyte Zitron et al. [82] (2005)
NeohesperidinXenopus oocyte Zitron et al. [82] (2005)
QuercetinXenopus oocyte Zitron et al. [82] (2005)
HEK 293 12 μ M Sun et al. [88] (2017)
RutinXenopus oocyte Zitron et al. [82] (2005)
HEK 293 Sun et al. [88] (2017)
TaxifolinHEK 293 Sun et al. [88] (2017)
Taxifolin 3-O- β -D-glucopyranosideCHO Yun et al. [92] (2013)
TrimethylapigeninHEK 29318–32 μ M Liu et al. [38] (2012)
The other study showed the inhibitory effects of acacetin [42]. It was revealed that this flavonoid is not only able to block the Kv11.1 channel ( I C 50 32 μ M) but also suppress channel current through the recombinant human cardiac Kv7.1 with its regulatory subunit KCNE1, which plays a prominent role in the repolarization of cardiac action potential [93]. Surprisingly, this suppression does not induce QT prolongation syndrome.
Another study conducted by Kelemen and her collaborators [87] demonstrated that epigallocatechin-3-gallate, similarly to naringenin, is able to block the Kv11.1 channel both in the HEK 293 cells ( I C 50 = 6 μ M ) and in Xenopus oocytes ( I C 50 = 20.5 μ M ) . Nevertheless, in contrast to naringenin, the action of EGCG is reversible. Moreover, Kelemen’s research group showed that the inhibitory effects of epigallocatechin-3-gallate are slow and do not disappear completely after a wash-out, which suggests the long-term effect of this compound on channel gating. These results did not find confirmation in the study conducted by Kang et al. [44]. In contrast to Kelemen et al., they observed only mild inhibitory effects of this flavonoid acting on the CHO cells.
In the article [85], Zhang et al. showed that genistein can block the hERG channel in a reversible manner. The authors concluded that this inhibition is probably not direct and mediated by the protein tyrosine kinase mechanism. In that work, daidzein, which is a tyrosine kinase-inactive analog of genistein, turned out to be a substantially less potent inhibitor of the Kv11.1 channel.
Recently, Sun et al. [88] published their results concerning the impact of an ensemble of different flavonoids on the Kv11.1 channel expressed in HEK 293 cells. They found that the strongest inhibitory effects are exerted by quercetin ( I C 50 = 12 μ M ) and fisetin ( I C 50 = 38 μ M ). Luteolin turned out to be a little less potent with half-blocking concentration I C 50 > 100 μ M . The weak inhibitory effects were observed for other analyzed compounds, such as galangin, kaempferol, and isorhamnetin.
Du et al. [86] observed that the extract of Galenia africana L. (Aizoaceae) stem and leaves enables effective inactivation of the hERG channel. It turned out that this extract is more potent than its constituents 7,8-methylenedioxyflavone and 7,8-dimethoxyflavone applied alone. Thus, the authors concluded that this inhibitory effect may stem from some synergistic interaction action between several components of the extract.
In [38], Liu and collaborators studied the impact of the methylated derivative of apigenin, trimethylapigenin, on the Kv11.1 channels. As it turned out, in contrast to the apigenin itself [82], trimethylapigenin was able to suppress the activity of Kv11.1 channel in a fully reversible, concentration-dependent manner with I C 50 18 μ M .
Another study carried out by Yun et al. [92] confirmed that also taxifolin 3-O-beta-D-glucopyranoside is able to effectively block the hERG channels in CHO cells. However, such inhibitory effects were not observed in terms of the administration of the taxifolin itself [90]. In addition, liquritigenin is able to inhibit the hERG channel at the moderate level ( I C 50 53 μ M ) by binding to the open state of the channel [91].

2.6. Further Kv Channels

Much less attention is paid in the literature to other types of Kv channels in the context of their modulation by the flavonoids. Let us provide the available information.
In [94], it was reported that the chronic administration of hesperetin was able to increase the expression of Kv1.2 channels in coronary arterial smooth muscle cells of diabetic rats. The authors conclude that since the expression of these ion channels is lowered in the diabetic rats, hesperetin can be considered a promising therapeutic agent in the treatment of coronary arterial dysfunction resulting from diabetes.
The flavonoids also influence the Kv7.1 channel, which contributes to the regulation of the repolarization phase of the cardiac action potential. Puerarin is an isoflavonoid found in the root of Pueraria Lobata, which is known from its anti-inflammatory, immunomodulatory, anti-cancer and cardioprotective properties [95]. In [43], it was demonstrated that the isoflavone, puerarin, effectively downregulates the channel activity via direct interaction with the channel protein. It was reported that this inhibitory effect (along with the blockade of the slow delayed rectifier current I K S ) contributed to the prolongation of action potential duration, which can be beneficial in the case of treatment of cardiovascular diseases. It turns out that also naringenin exerts an inhibitory effect on this channel, with a mild impact of this flavonoid on the I K S current [45]. Eventually, the studies performed by Kang et al. [44] revealed that also (−)-epigallocatechin-3-gallate [44] is a potent inhibitor of the Kv1.7 ion channel.
Quite recently, it has been discovered that another flavonoid, procyanidin B1, a natural compound extracted from the grape seed, is a potent inhibitor of the Kv10.1 channel ( I C 50 = 10 μ M ), which is overexpressed in some tumors [96]. According to this work, targeting the Kv10.1 channel by procyanidin B1 can inhibit the proliferation of cancerous cells. Consequently, this compound is a promising agent for cancer treatment.

3. Calcium-Activated Channels (KCa)

3.1. BK Channel

The large-conductance voltage- and C a 2 + -activated channels (BK) are ubiquitously expressed K + channels being characterized by a large single-channel conductance (150–300 pS) [97]. They are considered important drug targets due to their important roles in many physiological processes, such as neural transmission, hearing, endocrine secretion, and smooth muscle contraction [98]. In addition, the mitochondrial BK channel variants (mitoBK) received great scientific interest in terms of the possibilities of their chemical modulation because of the involvement of these channels in the regulation of metabolism, including ATP synthesis as well as the pro-life and pro-death processes [99,100,101].
The impact of flavonoids on the functioning of the BK channels was extensively studied in recent years. The main inferences from those investigations are outlined in Figure 1.
The activating effects were observed for the plasma membrane and mitochondrial BK channel variants in terms of the administration of a citrus flavanone, naringenin, in many different cell types [102,103,104,105,106,107,108,109,110]. The binding site for naringenin coordination seems to be located within the α subunits of the channel [103], probably within the gating ring. Naringenin coordination exerts the discernible effects on gating dynamics from the other stimuli [111] with no or negligible impact of auxiliary regulating β and γ subunits on naringenin binding. For this sake, it is anticipated that naringenin can be considered a general BK channel activator affecting all (or most) existing channel isoforms. According to [105], naringin exerts similar effects to naringenin on the BK channels. Another flavanone, dioclein, has been demonstrated to impose vasorelaxant effects, which can be, at least partially, explained by the activation of the BK channels and subsequent membrane hyperpolarization [112]. Among the flavanones, hesperedin also gained scientific interest as a modulator of C a 2 + -dependent channels. Namely, according to the in vitro study on the electrical activity of rat hippocampal cells [113], modulation of the BK channels is responsible for the anticonvulsive effects of hesperidin and its aglycone hesperetin.
Well-pronounced activating effects are reported in terms of the coordination of quercetin by the plasma-membrane BK channels exhibited in human bladder cancer cells, murine smooth muscles (ileal myocytes) and rat coronary smooth muscle cells [114,115,116]. In turn, the quercetin-related activation of the mitochondrial BK channel protein (mitoBK) was presented in human endothelial cell line EA.hy926 [117,118]. The physiological meaning of the quercetin-mediated BK and mitoBK channel activation is mainly associated with supporting the cell’s response to the oxidative stress [115] and cytoprotection [117]. Among other flavonols, the effective BK-opener profile was also shown for kaempferol in the work of Li et al. [119] in Xenopus oocytes injected with the mSlo gene. Moreover, the open-reinforcing impact of this flavonoid on the BK channel is responsible for relaxation of the rat pulmonary artery through the membrane hyperpolarization [120], which stays in agreement with the results obtained in [121]. In that work, the vasodilatory effects of kaempferol were related to its ability to stimulate the BK channels in human umbilical vein endothelial cells. The kaempferol-enhanced endothelium-dependent relaxation was observed in the porcine coronary artery, and it was also mediated by the activation of the BK channels [122].
Recent studies showed that luteolin acts as the mitoBK channel activator in cardiomyocytes and endothelial cells [123]. Luteolin-mediated mitoBK channel activation can contribute to the well-documented cardioprotective effects of this flavonoid. In contrast, another representative of flavones, nobiletin, can be considered a BK channel inhibitor, which acts in a voltage- and C a 2 + -dependent manner [124]. It is interesting that the presence of different regulating β subunits affects the efficacy of nobiletin action, which suggests that the inhibitory effects of this flavone can exhibit tissue selectivity, since the accessory BK channel subunits are frequently expressed in a tissue-dependent manner.
Considering the impact of other flavones on BK channels, in the studies on the relaxant effects of baicalein on tracheal smooth muscle [125], the authors anticipate that the molecular mechanism of bronchodilation induced by this substance incorporates the increase in the frequency of BK channels’ opening. Additionally, another flavone, apigenin, is an effective BK channel activator, according to the results obtained in [119] on Xenopus oocytes transfected with the mSlo gene. Among this group of flavonoids, the endothelium-dependent morelloflavone-induced vasorelaxation was observed in the experiments on isolated rat thoracic aorta, which was precontracted with norepinephrine [126]. This process partly involved the BK channels activation (and the opening of ATP-sensitive K + channels, as discussed in the next section).
Stimulation of the BK channels by isoflavonoid genistein resulted in both hampering and enhancing effects for the transport capabilities of the channel depending on the cell types under study. In the in vitro experiments on the selected molecular aspects of atherosclerosis described in [127], genistein inhibited the BK channels in the vascular smooth muscle cells stimulated by oxidized low-density lipoprotein. This kind of modulation resulted in the suppression of proliferation of those cells. The decrease of the BK channel activity in the presence of genistein (and genistein in combination with M g 2 + ) was also observed in vascular smooth muscle cells in the rat model of hypertension [128] as well as in a similar rat model of hemorrhagic shock [129]. On the other hand, the voltage-dependent BK current was increased by this isoflavonoid in the case of the HEK 293 cells transfected by BK channels (i.e., human BK channel α and the β 1 pcDNA3.1 plasmids) [130]. Moreover, the genistein-induced BK channel-activating effects were observed in bovine trabecular meshwork cells [131].
Another isoflavonoid, daidzein, acts as a concentration-dependent activator of the BK channel in complex with β 1 subunit according to the results obtained in [132] in experiments carried out on rat cerebral basilar artery smooth muscle cells. Additionally, in the studies of vasorelaxation induced by genistein and daidzein in noradrenaline and KCl precontracted rat mesenteric artery preparations [133], it turned out that iberiotoxin ( c = 1–10 nM) and charybdotoxin ( c = 30 nM), being well-known antagonists of the BK channel, inhibited relaxation. These observations suggest activating effects of daidzein and genistein for BK channels in the case of the analyzed cells. The BK-stimulating effects of daidzein were confirmed in the investigations on Xenopus oocytes expressing mSlo [90]. However, better-pronounced effects were established in the case of channel stimulation by its analog, puerarin. Moreover, the authors observed the highest channel-activating potency of puerarin, when the BK channel was transfected in a form of mSlo–h β 1 complex, and this open-reinforcing effect can underlie the puerarin-mediated vasodilation [90]. Puerarin can also act as a mitochondrial BK channel modulator. According to the studies carried out on rat cardiomyocytes [134], pretreatment of the investigated cells with puerarin at c = 0.24 mM for 5 min increased the cell viability against H 2 O 2 -stress. Further analysis indicated that the protection of cardiomyocytes against H 2 O 2 -stress by puerarin is mediated by the activation of mitochondrial BK channels. The mitoBK channel activation by puerarin was also observed in [135]. In that study, puerarin at c = 0.24 mM protected rat myocardial cells from hypoxia/reoxygenation damage by enhancing the mitochondrial K + transport via mitoBK and mitoKATP channels (as discussed in the next section).
The open state probability of the BK channels in myelinated nerve fibers of Xenopus laevis was greatly increased by external phloretin ( c = 10–200 μ M). The analysis of the patch-clamp recordings of the BK channels stimulated by this chalcone showed that the open dwell times were prolonged and closed dwell times were shortened in relation to control data [136]. The action of phloretin, as a BK channel opener, was confirmed in [137,138], where the authors studied heterologously expressed BK channels composed of human α subunits in different concentrations of calcium ions and over a wide range of membrane potentials. Another chalcone, nothofagin, elicited endothelium-dependent vasodilation in the perfused rat kidney, and this effect is mediated by activation of the large-conductance potassium channels [139].
An alkaloid, berberine, is conditionally considered by some authors as an ‘isoquinoline flavonoid’ [140]. One of the cricial factors responsible for its biological meaning is BK channel modulation. According to the studies conducted on streptozotocin-induced diabetic rats [141], the chronic administration of berberine (100 mg/kg/day) can lower blood glucose level, reduce blood pressure and improve vasodilation. The important mechanism underlying that finding is that berberine markedly increased the open state probability and expression level of BK channels coordinated with β 1-subunits in cerebral vascular smooth muscle cells isolated from diabetic rats or when exposed to hyperglycemia condition. Moreover, according to the research on Sanoshashinto, which is a classical prescription in China and Japan against hypertension, its key components, berberine and baicalin, are suggested to be responsible for the observed vasorelaxant effects [142]. It is hypothesized that the biological consequences of administration of these substances stem from the opening of the BK channels together with the activation of other pathways (the NO/cGMP and the DAG/PKC/CPI-17 pathway).
Another flavonoid that causes vasodilation is rottlerin. Such a physiological effect is mediated by rottlerin-induced BK channel activation, according to the rat and mouse models of cardioplegic arrest and reperfusion [143]. The open-reinforcing effect of rottlerin (in micromolar concentration) on BK channels was also observed in the studies performed on murine tracheal smooth muscles [144] (where it supported airway smooth muscle relaxation). The rottlerin-mediated BK channel activation was also detected in human hepatic stellate cells, where it was important for the liver profibrotic signaling pathways [145].
The summary of flavonoid modulation of BK channels activity is presented in Table 3.

3.2. IK and SK Channels

The small- and intermediate-conductance C a 2 + -dependent K + channels are not as extensively studied in terms of their effective stimulation by flavonoids as their large-conductance counterparts. Nevertheless, some reports emphasize the involvement of IK and SK channels in shaping the physiological response to flavonoid stimulation.
First, the molecular mechanism of vasodilation in rat aorta induced by quercetin is suggested to incorporate mainly activation of the SK channels, according to the works [146,147,148]. The SK channels play an important role in the cardioprotective effects of prolonged administration of an extract from leaves of Croton urucurana Baill. (Euphorbiaceae), which is popularly known as ‘sangue de dragão’, according to the spontaneously hypertensive rat model [149]. Flavonoids (including rutin, isoquercetin, kaempferol, vitexin) are the key bioactive substances in this extract [150]. The work [151] demonstrates that the vascular relaxation of rat aortic rings caused by a crude hydroalcoholic extract from Polygala paniculata (rich in rutin) involves the nitric oxide/guanylate cyclase pathway and subsequent opening of IK and BK channels. These effects are, however, larger in vitro than in vivo.
Among flavones, acacetin is an SK channel blocker, as confirmed in investigations on the small-conductance C a 2 + -dependent K + channels expressed in HEK 293 cells [152]. These studies evidenced that acacetin inhibited three subtypes of the SK channels (SK1, SK2, SK3) in a concentration-dependent manner with I C 50 of 12.4 μ M for SK1, I C 50 = 10.8 μ M for SK2, and I C 50 = 11.6 μ M for SK3. The former experiments performed on a canine model (using isolated canine left atrium) showed that blockade of the SK channels by acacetin likely contributes to its anti-atrial fibrillation property [153]. Another flavone, isovitexin, obtained from the extract of Luehea divaricata Mart. regulates mesenteric arteriolar tone due to the activation of the SK channels and the Kir6.1 ATP-sensitive K + channels [154].
The studies on the vasorelaxant effects of genistein and daidzein administration [133] suggest that the activity of the SK channels can be modulated by these flavonoids and mediate the observed biological effects. The genistein- and daidzein-induced relaxation of rat noradrenaline precontracted arterial rings was decreased by apamin ( c = 0.1–0.3 μ M), being an antagonist of the SK channels.
Some studies on the effects of flavonoids on the activity of the C a 2 + -gated channels unraveled that although a given flavonoid modulates the activity of the BK channels, it does not interact with either SK or IK channels. Such an observation was made in the work of Xu et al. [120,121,122], where kaempferol had no effect on the SK and IK channels.
In turn, the studies of the vasodilatory properties of nothofagin [139] exclude the involvement of the SK channels in the mediation of that effect and indicate the main role of the BK channels. Nevertheless, a hypothesis on a potential additional contribution of the IK channels in the observed nothofagin-induced vasodilation cannot be rejected.

4. Inward Rectifying Potassium Channels (Kir)

The Inward Rectifying Potassium Channels ( K i r ) belong to one of the structurally simplest ion channels group containing four identical subunits, each containing two membrane-spanning alpha helices. These channels allow ions to be transported more effectively into than out of the cell. They are responsible for the regulation of resting membrane potential. Thus, their function is mostly related to the modulation of cardiac and neural cells activity, insulin secretion, or epithelial K + transport [155]. The Kir channels are expressed in many cell types: myocytes, neurons, blood cells, endothelial, glial cells, or oocytes [156]. The classification of the Kir channels family covers the groups Kir1–Kir7 together with their respective subgroups. Among the Kir channels, one can also distinguish the adenosine triphosphate (ATP)-dependent K + channels (KATP, Kir6) and the G-protein regulated K + channels (GIRK, Kir3). The structure of Kir channels lacks a proper voltage-sensing domain. Nevertheless, some representatives of the Kir family exert a bit stronger ‘’voltage dependence” than the others. In that aspect, Kir 2 channels, which are strongly rectifying ones (and consequently more sensitive to extracellular K + ), deserve to be distinguished.
The Kir channels can interact with a wide range of molecules, including flavonoids. Below, we shortly characterize the effects of flavonoid administration on the activity of inward rectifier potassium channels.
The literature indicates that flavones can affect the Kir channels’ activity. Jiao et al. [157] proved that flavones from rhododendron can stimulate the opening of ATP-dependent Kir channels in rat cardiomyocytes, which is related to the cardioprotective effects of this group of flavonoids. Another example of flavone being important in the context of Kir channel stimulation is luteolin. Li et al. [158] proved the positive impact of this compound for Kir channels present in rat coronary arterial smooth muscle cells, which was associated with inhibition of the process of vasoconstriction.
The flavonols represented by quercetin and rutin exert significant impact on Kir channels. Trezza et al. [159] examined the both modulators and 5-hydroxyflavone in the context of their possible impact on the ATP-sensitive Kir6.1 channel. They compared the experimental results of channels from Rat norvegicus aorta cells with molecular dynamics and docking calculations. All the compared results suggested that there was no effect on Kir6.1 caused by rutin, and significant downregulation in the case of quercetin and 5-hydroxyflavone, but only in the case of the closed channel conformation. The cardioprotective effect of flavonoids on rat myocytes through the regulation of mitochondrial ATP-sensitive potassium channels activity was also shown recently by Rameshrad et al. [160]. This research group studied a flavonol, morin, and postulated that its antioxidative effects are mediated by mitochondrial ATP-dependent potassium channels. The activity of Kir6.1 can be upregulated in the presence of isovitexin, which is obtained from the extract of Luehea divaricata Mart., according to [154]. This effect supports the regulation of mesenteric arteriolar tone.
In [161], the authors studied the physiological effects of the administration of Rooibos tea (Aspalathus linearis) as well as its pure flavonoid components: chrysoeriol, vitexin, and orientin. These substances were anticipated to mitigate hyperactive gastrointestinal disorders as well as exert health-beneficial effects in cardiovascular and respiratory diseases. With this aim, the research was conducted on fresh preparations of rabbit jejunum and aortic rings, guinea-pig trachea, and right atria. The main conclusion referred to the selective bronchodilator effect of Rooibos tea, which turned out to be mediated through KATP channel activation by chrysoeriol. This flavonoid also induced KATP-mediated relaxations of precontracted jejunum and aortic preparations by low ( c = 25 mM) K + without any effect on high ( c = 80 mM) K + -induced contractions. Vitexin inhibited low K + -induced contractions in jejunum and trachea, while orientin exerted only the relaxation effect in jejunum.
The in vitro studies on Xenopus oocytes expressing Kir6.2 channels showed that (−)-epigallocatechin-3-gallate and (−)-epicatechin-3-gallate (ECG) inhibit the activity of these ATP-sensitive potassium channels [162]. It turned out that ECG is three times more effective than EGCG. Two other compounds, (−)-epicatechin and (−)-epigallocatechin, did not affect the channel activity. Because the authors introduce structural modifications of the channel, in the conclusion, they formulate some hypotheses about the possible binding sited for EGCT within the Kir6.2 protein structure. In the same work, the authors analyze also the effects of EGCG on insulin secretory responses to high glucose loading in an in vivo rat model (hampering).
Naringenin has proven anti-inflammatory and antioxidant properties, which can be partly associated with the regulation of the ATP-sensitive potassium channels. In [163], Pinho et al. characterized this modulator in the context of activation of the NO–cyclic GMP–PKG–ATP-sensitive K + channel pathway, which can be related to the reduction of oxidative stress and translates into a decrease of inflammatory pain in mice. Similar results were obtained by Manchope et al. [164], where the activation of the same pathway leads to the reduction of the nociceptor hyperpolarization, and, in consequence, to the inhibition of its neuronal transmission. Another article that takes into consideration the function of ATP-sensitive channels in conjunction with the opioid receptors and the action of naringenin was written by [165], where the L-arginine/NO/cGMP/KATP pathway was analyzed. According to the studies performed on a rat model of ischemia–reperfusion (I–R) injury by Meng et al. [166], naringenin at a concentration above 2.5 μ M activates KATP channels in both the plasma membrane and the mitochondria. In turn, the KATP channels’ activation contributes to the cardioprotective properties of naringenin.
Well-pronounced effects on the Kir channels (especially KATP channels) are also documented for a natural alkaloid, berberine (BBR), which is conditionally classified as an ’isoquinoline flavonoid’. BBR is frequently used in the Chinese and East Asian medicines [140]. Hua et al. confirmed the inhibition effect of BBR on ATP-sensitive channels [167]. The authors postulated that the anti-arrhythmic and antidiabetic properties of berberine are related to the inhibition of potassium channels. The inhibitory effects of berberine were also investigated by Wang et al. [168]. The authors characterized a similar anti-arrhythmic impact of BBR manifested by the reduction of action potential duration and the effective refractory period of ischemia. In contrast to BBR, a flavonoid from the anthocyanins-cyanidin caused the upreguation of Kir6.2 genes, which have potential implication in glucose sensitivity and its homeostasis [169]. Another flavonoid often used in Chinese and Japanese natural medicine is baicalein. The positive health aspects of baicalein is described in the context of potassium channels modulation by Saadat et al. [125]. In this work, the authors postulated that baicalein is an activator of ATP-dependent potassium channels in rat tracheal smooth muscle and is involved in bronchodilation through the promotion of the K + channel opening. Ribeiro et al. [170] suggest that the activation of the ATP-sensitive K + channels by baicalein can underlie the gastroprotective properties of this flavonoid.
A group of flavonoids with a strong influence on the Kir channels is isoflavonoids. Among them, genistein is considered the most common modulator of inwardly rectifying potassium channels. This compound exhibited a typical Kir channels inhibiting profile in several studies. Zhao et al. [171] described the molecular character of this inhibition. For the Kir2.3 channel, it was proved that the key protein regions responsible for the genistein-related inhibition are transmembrane domains and the pore. Ko et al. [172] showed that genistein blockade depends on the mode of the channel activity—the modulator did not exert any effect on the steady-state activation or inactivation of Kir channels. In the work written by Okamoto et al. [173], it was shown that the reduction of the Kir current induced by genistein entailed the depolarization of membrane of rat osteoclast, and in final effect, it caused an elevation of C a 2 + and inhibition of osteoclastic bone resorption. The earlier work by Okata et al. [174] had suggested that the tyrosine kinase may be involved in the inhibitory character of the impact of genistein on ATP-dependent channels.
Another isoflavonoid, puerarin, exerts an activating effect on the mitochondrial K + ATP-regulated channels (mitoKATP), according to the results presented in [135]. That study concluded that the mitoKATP channel activation participated in the cardioprotection by puerarin. The activation of mitochondrial KATP channels also plays a crucial role in shaping the cardioprotective effects exerted by other flavonoids [175]. Among them, six natural compounds should be mentioned: (−)-epigallocatechin-3-gallate [176], theaflavin [177], proanthocyanidins [178], genistein [179], baicalein [180], and morin[160].
As one can see, flavonoids can be considered KATP channel modulators. The summary of flavonoid modulation of ATP-sensitive Kir channels is outlined in Table 4.
Yow et al. [182] characterize the impact of naringin as a direct activator of the G protein-coupled Kir channel, which is important in CNS control and heart rate regulation. A flavanone, hesperidin, interacts with the G protein-activated GIRK1 and GIRK2 channels and causes their inhibition, according to the results presented in [183]. It turns out that hesperidin inhibits GIRK1 and GIRK2 currents through binding to the μ -opioid receptor, and it participates in the anti-depressant and antinociceptive activities of hesperidin. The GIRK current may be also inhibited by eriodictyol, from a flavanone group, which occurs in citrus fruits and Chinese herbs. The inhibition character of this flavonoid on GIRK channels was documented in the work of Hammadi et al. [184]. The impact of flavonoids on the G-protein activated Kirs is summarized in Table 5.

5. Two-Pore Domain Potassium Channels (K2P)

The two-pore domain potassium channels are widely distributed in excitable and non-excitable cells and are responsible for the background potassium conductance [185,186]. They are emerging drug targets in case of a.o. cardiovascular and neurological diseases [187,188,189,190]. The K2P channels are a family of 15 K + channel subtypes, including the TWIK channels (named as an acronym for Tandem of pore domains in a Weak Inward rectifying K + channels) K 2 p 1.1 , K 2 p 6.1 and K 2 p 7.1 , TREK channels (TWIK-related K + channels): K 2 p 2.1 and K 2 p 10.1 , TRAAK channel (TWIK-related arachidonic acid-activated potassium channel) belonging to the TREK subgroup: K 2 p 4.1 , TASK channels (TWIK-related acid-sensitive K + channel channels): K 2 p 3.1 , K 2 p 5.1 , K 2 p 9.1 and K 2 p 15.1 , THIK channels (tandem pore domain halothane-inhibited K + channels): K 2 p 12.1 and K 2 p 13.1 , TALK channels (TWIK-related alkaline pH-activated K + channels): K 2 p 16.1 and K 2 p 16.1 , and TRIK channel (TWIK-related spinal cord K + channel) K 2 p 18.1 .
Among the TREK subgroup of K2P channels, TREK-1 ( K 2 p 2.1 ) and TRAAK ( K 2 p 4.1 ) are mainly expressed in the central nervous system (CNS), and TREK-2 ( K 2 p 10.1 ) is expressed in both CNS and peripheral tissues [191,192]. TREK channels are activated by several stimuli, including biomolecules (e.g., riluzole, nitrous oxide, polyunsaturated fatty acids, and lysophospholipids). These modulators can contribute to the opening of TREKs under pathological conditions. Considering the effects of flavonoids’ administration, the neuroprotective properties of quercetin were demonstrated in [193]. In that study, the mice manic model was induced by i.p. injection of D-amphetamine, and quercetin suppressed the neural excitability of prefrontal cortex pyramidal neurons. This effect was mediated by enhancing current flow through TREK-1 channels, which decreased membrane resistance. In [194], the authors demonstrate that baicalein and wogonin increased the open state probability of TREK-2 channels in a dose-dependent manner (in the range from 0 to 100 μ M, at which the maximal channel-activating effect was observed), leaving the single-channel conductance and mean open dwell-time unchanged. These studies were carried out on the COS-7 cells (African green monkey kidney fibroblast-like cell line) transfected with rat TREK-2. Since baicalein elicited a continuous channel-activating effect, while wogonin activated the TREK-2 channel transiently, it was anticipated that these flavones interact with the TREK-2 channel protein by different molecular mechanisms. Nevertheless, the TREK-2 modulation by wogonin and baicalein may exert beneficial effects in neuroprotection.
The studies on the possible impact of tyrosine kinase inhibitor, genistein, on the activity of the human TASK-1 ( K 2 p 3.1 ) channel expressed in Xenopus oocytes and Chinese hamster ovary cells (CHO) revealed the blocking effects ( I C 50 = 10.7 μ M in Xenopus oocytes and I C 50 = 12.3 μ M in CHO cells) [195,196]. These studies showed that an isoflavonoid, daidzein (at a concentration of 100 μ M), causes 18.2 ± 1.3% inhibition of human TASK-1 expressed in Xenopus oocytes [196]. Moreover, the TASK-3 ( K 2 p 9.1 ), THIK-1 ( K 2 p 13.1 ) and TWIK-2 ( K 2 p 6.1 ) currents also decreased in the presence of genistein [195,196], and the same relation for the TASK-2 ( K 2 p 5.1 ) activity was suggested in [197]. These observations allowed one to make the inference that inhibition of the K2P currents via biochemically induced changes in tyrosine kinase activity permits membrane potential depolarization and excitation.

6. Discussion

Flavonoids are widely known for their beneficial health effects, which involve complex biochemical interactions with specific molecular targets, including potassium channels. Due to the fact that these transport proteins play important roles in shaping cardiac action potential as well as smooth muscle tone, their stimulation by flavonoids yields vasorelaxant and cardioprotective effects [81,148,175,198,199,200]. Nevertheless, these effects are not the only examples of the K + channel-mediated physiological processes that are regulated by flavonoids, as presented in Figure 2.
In this work, we summarized the state of knowledge about flavonoids as modulators of particular subtypes of K + channels. It allowed us to point out the most promising natural substances for further research from the pharmacological point of view. Such future studies can include the analysis of their derivatives to develop novel substances which may exert better specificity and efficiency against particular channel proteins. However, this is a challenging task due to the complex mechanisms of flavonoid interactions with channel proteins (direct and/or indirect via second-messenger proteins or the changes of membrane properties) as well as multiple possibilities of the structure–function modification. As an example, let us refer to the studies on quercetin being an effective mitoBK channel activator and its analog isorhamnetin which turned out to not affect the mitochondrial BK channel activity [118]. On the other hand, the relatively weak inhibitory effect of the apigenin on the Kv1.5 channel is improved when the methylated derivatives of this flavonoid (dimethylapigenin or trimethylapigenin) are introduced at the same concentrations [68]. Furthermore, the enhanced inhibition of the Kv 1.3 channel was observed when the active compounds possess a prenyl group in its structure in comparison to their non-prenylated counterparts (such as 8-prenylnaringenin, isoxanthohumol in relation to their non-prenylated analogs naringenin, genistein) [31,32].
As can be observed in Table 1, Table 2, Table 3, Table 4 and Table 5, some flavonoids (e.g., naringenin, quercetin, genistein) have a wide spectrum of molecular targets within the family of K + channels. It suggests the existence of multiple possible binding sites for these biomolecules within channel protein structures or a number of second-messenger molecules in case of indirect mechanisms or the relatively large effects exerted by the physicochemical modulation of membrane properties by flavonoids, which accounts for the flavonoid–channel interactions. According to the literature, the last factor, i.e., the interactions of flavonoids with the membrane, and the consequent changes of the membrane composition, packaging, fluidity, permeability and interactions of its lipid components with the membrane proteins [201,202,203] significantly modify membrane-mediated cell signaling cascades. Thus, it is partly responsible for the pharmacological activities of flavonoids, including its anti-tumor, anti-microbial and anti-oxidant properties [204,205,206,207]. The effects exerted by flavonoids on the biological membranes are mainly associated with their planar structure and lipophilicity, which are dependent on, among others, the number and position of hydroxyl groups [201,203]. Relatively hydrophobic flavonoids (such as flavones) are incorporated into the interior of the lipid bilayers, and they increase the ordering and dynamics within the internal (fatty) part of the membrane. In contrast, more hydrophilic flavonoids (such as flavonols) interact primarily with the membrane surface. The localization and strength of the flavonoids–membrane interactions affect the functioning of integral membrane proteins (including ion channels) and modulate their structure and function [208,209,210,211]. That is because the membrane proteins strongly interact with their lipid surroundings. They are not rigid entities, but to ensure a good hydrophobic matching to the adjacent lipid bilayer, they undergo structural deformations. Consequently, due to such deformations, any change of the membrane properties can result in the allosteric modulation of ion channels functioning. From this perspective, considering the molecular mechanisms of flavonoid–channel interactions, there exists a strong correlation between the structure and the molecular activity of a flavonoid. It stems from the synergistic effects of the effective and specific binding of a given flavonoid to a particular target protein (e.g., ion channel) and the additional indirect interactions mediated by the membrane. As an example, the differences in lipid composition in cell membrane and mitochondrial membranes are anticipated to contribute to the possible quantitative differences in the flavonoid-mediated activation levels of plasma-membrane and mitochondrial variants of potassium channels (e.g., BK/mitoBK channels) [203]. Analogous effects can make a contribution to the quantitative differences between the outcomes of channel activation by a given flavonoid in different cell types.
Due to the existing structural and mechanistic differences implying other binding sites for flavonoids or various paths of indirect interactions, even if a given flavonoid can regulate different channel types, the directions of these modulations can be completely different. For instance, quercetin is an activator of the BK channels (from plasma membrane and their mitochondrial analogues) [114,115,116,117,118], but it has an inhibitory effect on the Kir6.1 channels [159]. Analogously, naringenin stabilizes the open state of the BK and mitoBK channels [102,103,104,105,106,107,108,109,110]. At the same time, it has a multichannel inhibitory profile against hERG, Kir2.1, Kv7.1, and Kv4.3 channels [45].
For some flavonoids, many details of the molecular mechanism of their specific interactions with channel proteins become unraveled, as in the case of naringenin and quercetin coordination to the plasma membrane/mitochondrial BK channels [103,111,118] or Kir6.1 modulation caused by quercetin and 5-hydroxyflavone [159]. Nevertheless, a clear picture of the possible direct or indirect interactions between most flavonoids and channel types remains unknown. Therefore, it can become a field of exploitation for both experimental and in silico studies.
Apart from the reports mentioned before within this review, which precisely describe the effects of flavonoid administration on particular subtypes of potassium channels, in the literature, one can also encounter the ones that provide only general information about the involvement of K + channels in mediating the flavonoid-induced physiological effect. In such studies, either a non-specific K + channel blocker (tetraethylammoniumchloride) or combinations of different blockers were applied in the experimental work. Let us provide a few examples. Sinensetin from Orthosiphon stamineus Benth. (Lambiaceae) results in vasorelaxation. In turn, the strong inhibition of the vasorelaxant effects elicited by this flavonoid was observed in terms of administration of the potassium channel antagonists, which suggests the employment of different pathways involving the Kir, KCa and Kv channels in the considered phenomenon [212]. Analogous vasorelaxant effects mediated by K + channels were observed for the rat aortic rings treated with luteolin [213]. The investigations on biochemical responses in terms of the cellular stress induced by apigenin isolated from Aster yomena in Candida albicans [214] indicated that apigenin induced ion channel-mediated potassium leakage. In turn, the important role of potassium channels in the attenuation of neurotoxic mitochondrial calcium overload by a Citrus polymethoxylated flavone, nobiletin, was suggested in [215]. Another study, which focuses on the antidepressant-like effect of hesperidin in a Tail Suspension Test in mice, demonstrated the contribution of the KATP and KCa channels in this phenomenon [216]. Moreover, hesperidin and its aglycone, hesperetin, are associated with beneficial outcomes for human health, such as prevention of cancer and counteracting cardiovascular diseases, which is partly due to the K + channel modulation according to [217]. Thus, to extract the exact information about the particular types of potassium channels affected by the mentioned flavonoids, further research is needed.
Considering the other interesting directions for further research, we are convinced that a thorough extended analysis of the physiological pathways (possibly, partly mediated by the K + channels) responsible for the antidiabetic, anti-inflammatory, and anti-cancerogenic effects of the flavonoids’ administration could be recommended. An additional interesting approach is to synthesize hybrid molecules made of two different substances with synergistic action in modulating potassium channels. As an example, this approach has been successfully realized in the synthesis of celecoxib with the flavonoid combrestatin A-4 with the goal to improve the anti-inflammatory properties of both substances[218]. This could yield valuable contributions as a response to the popular and challenging public health problems worldwide.
There are already some reports on the regulation of insulin homeostasis and metabolic processes by flavonoids, as summarized in [219]. In this work, we have also mentioned the studies which associate the antidiabetic effects of berberine with KATP channels’ inhibition [167]. Moreover, according to [141], this substance counteracts the diabetes-related vascular complications via BK channel activation in cerebral smooth muscle cells. In the antidiabetic context, naringenin (repeatedly mentioned in this review as a K + channel modulator) also deserves particular attention [220,221]. The metabolic metabolic diseases including obesity, metabolic syndrome and type 2 diabetes (T2D) are gathered by the excess adiposity, which sustains a state of chronic low-grade inflammation. In turn, chronic inflammation is an important factor contributing to DNA damage and can lead to cancer [222,223]. From this perspective, the administration of flavonoids, which exhibit a relatively wide spectrum of beneficial effects, such as promoting anti-oxidation (as mentioned in case of the, e.g., mitochondrial BK channel modulation) and immunosuppression (as discussed in Section 2), seems reasonable especially for the high-risk populations, e.g., suffering T2D [220,221] or endocrinopathies [224,225,226].
Another interesting aspect of flavonoid delivery is the alleviation of the effects of hormonal imbalance [224,225,226]. For instance, genistein and daidzein are estrogen-like compounds, xenoestrogens, that can bind competitively to estrogen receptors. Thus, they are considered an alternative to hormone replacement therapy in postmenopausal women or patients with some ovary dysfunctions [227]. What is worth mentioning is that some of their beneficial effects are mediated by K + channels. In general, the usage of phytoestrogens can be also recommended from the perspective of the possible neuroprotection [228], prevention of cancer [229], and atherosclerosis [133]. Still, the determination of the direct effects of flavonoid xenoestrogens on the endocrine system can be an interesting subject for further research.
To sum up, the molecular aspects of the prophylactic and therapeutic effects of flavonoids can become a promising field of exploitation for further biochemical and pharmacological investigations. This direction of research could provide a scientific justification for the flavonoid supplementation with the aim of prevention and treatment of popular diseases, where prolonged conventional therapy can be burdensome for patients and exhibit side effects.

7. Conclusions

Flavonoids are a group of natural substances that can effectively interact and regulate the functioning of many potassium channel types, which has been outlined in this review. The modulation of K + channels by flavonoids and their derivatives, together with their physiological consequences, should be subject to further investigation as a promising approach to the prevention or treatment of, among others, cardiovascular and inflammatory diseases.

Author Contributions

Conceptualization, A.W.-J.; formal analysis, A.W.-J. and D.V.D.; investigation, M.R.-L., A.W.-J. and P.T.; resources, M.R.-L., A.W.-J. and P.T.; data curation, M.R.-L., A.W.-J. and P.T.; writing—original draft preparation, M.R.-L., A.W.-J. and P.T.; writing—review and editing, M.R.-L., A.W.-J., P.T. and D.V.D.; visualization, A.W.-J. and P.T.; supervision, D.V.D.; project administration, M.R.-L. and A.W.-J.; funding acquisition, M.R.-L. and A.W.-J. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Łukasiewicz Research Network–Institute of Medical Technology and Equipment as part of a subsidy from the Ministry of Education and Science to M.R.-L. and the Silesian University of Technology grant for young researchers No. 04/040/BKM22/0216 to A.W.-J. (statute project). The funding organizations had no role in study conceptualization, analysis, decision to publish, or preparation of the manuscript.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
AFatrial fibrillation
APaction potential
BKbig-conductance (large-conductance) C a 2 + -dependent potassium channels
CNScentral nervous system
E C 50 half maximal effective concentration
EGCG(−)-epigallocatechin-3-gallate
I C 50 inhibitory concentration 50%
IKintermediate-conductance C a 2 + -dependent potassium channels
I K u r ultra-rapid delayed rectifier current
K2Ptwo-pore domain potassium channels
KCa C a 2 + -regulated potassium channels
Kvvoltage-regulated (voltage-gated) potassium channels
Kirinward rectifier potassium channels
mitoKATPmitochondrial K + ATP-regulated channels
mitoBKmitochondrial BK channels
PHpulmonary hypertension
SKsmall-conductance C a 2 + -dependent potassium channels
T2Dtype 2 diabetes
TMDstransmembrane domains
TALKTWIK-related alkaline pH-activated K + channels
TASKTWIK-related acid-sensitive K + channel channels
THIKtandem pore domain halothane-inhibited K + channels
TRAAKTWIK-related arachidonic acid-activated potassium channel
TREKTWIK-related K + channels
TRIKTWIK-related spinal cord K + channel
TWIKtandem of pore domains in a weak inward rectifying K + channels

References

  1. Hille, B. Ionic channels in excitable membranes. Current problems and biophysical approaches. Biophys. J. 1978, 22, 283–294. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Coetzee, W.A.; Amarillo, Y.; Chiu, J.; Chow, A.; Lau, D.; McCormack, T.; Morena, H.; Nadal, M.S.; Ozaita, A.; Pountney, D.; et al. Molecular diversity of K+ channels. Ann. N. Y. Acad. Sci. 1999, 868, 233–255. [Google Scholar] [CrossRef] [PubMed]
  3. Miller, C. An overview of the potassium channel family. Genome Biol. 2000, 1, reviews0004.1. [Google Scholar] [CrossRef] [PubMed]
  4. MacKinnon, R. Potassium channels. FEBS Lett. 2003, 555, 62–65. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Kuang, Q.; Purhonen, P.; Hebert, H. Structure of potassium channels. Cell. Mol. Life Sci. 2015, 72, 3677–3693. [Google Scholar] [CrossRef] [Green Version]
  6. Wulff, H.; Castle, N.A.; Pardo, L.A. Voltage-gated potassium channels as therapeutic targets. Nat. Rev. Drug Discov. 2009, 8, 982–1001. [Google Scholar] [CrossRef] [Green Version]
  7. Hutchings, C.J.; Colussi, P.; Clark, T.G. Ion channels as therapeutic antibody targets. Proc. Mabs. Taylor Fr. 2019, 11, 265–296. [Google Scholar] [CrossRef]
  8. Mathie, A.; Veale, E.L.; Cunningham, K.P.; Holden, R.G.; Wright, P.D. Two-pore domain potassium channels as drug targets: Anesthesia and beyond. Annu. Rev. Pharmacol. Toxicol. 2021, 61, 401–420. [Google Scholar] [CrossRef]
  9. Cui, M.; Cantwell, L.; Zorn, A.; Logothetis, D.E. Kir Channel Molecular Physiology, Pharmacology, and Therapeutic Implications. Pharmacol. Potassium Channels 2021, 267, 277–356. [Google Scholar]
  10. Dudem, S.; Sergeant, G.P.; Thornbury, K.D.; Hollywood, M.A. Calcium-activated K+ channels (KCa) and therapeutic implications. In Pharmacology of Potassium Channels; Springer: Berlin, Germany, 2021; pp. 379–416. [Google Scholar]
  11. Zúñiga, L.; Cayo, A.; González, W.; Vilos, C.; Zúñiga, R. Potassium Channels as a Target for Cancer Therapy: Current Perspectives. Oncotargets Ther. 2022, 15, 783–797. [Google Scholar] [CrossRef]
  12. Wulff, H.; Christophersen, P.; Colussi, P.; Chandy, K.G.; Yarov-Yarovoy, V. Antibodies and venom peptides: New modalities for ion channels. Nat. Rev. Drug Discov. 2019, 18, 339–357. [Google Scholar] [CrossRef] [PubMed]
  13. Adorisio, S.; Fierabracci, A.; Rossetto, A.; Muscari, I.; Nardicchi, V.; Liberati, A.M.; Riccardi, C.; Van Sung, T.; Thuy, T.T.; Delfino, D.V. Integration of traditional and western medicine in Vietnamese populations: A review of health perceptions and therapies. Nat. Prod. Commun. 2016, 11, 1409–1416. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Yuan, H.; Ma, Q.; Ye, L.; Piao, G. The traditional medicine and modern medicine from natural products. Molecules 2016, 21, 559. [Google Scholar] [CrossRef] [Green Version]
  15. Georgiev, M.I. From plants to pharmacy shelf: Natural products revival. Phytochem. Rev. 2016, 15, 511–513. [Google Scholar] [CrossRef] [Green Version]
  16. Gechev, T.S.; Hille, J.; Woerdenbag, H.J.; Benina, M.; Mehterov, N.; Toneva, V.; Fernie, A.R.; Mueller-Roeber, B. Natural products from resurrection plants: Potential for medical applications. Biotechnol. Adv. 2014, 32, 1091–1101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  17. Monjotin, N.; Amiot, M.J.; Fleurentin, J.; Morel, J.M.; Raynal, S. Clinical evidence of the benefits of phytonutrients in human healthcare. Nutrients 2022, 14, 1712. [Google Scholar] [CrossRef] [PubMed]
  18. Wang, Z.; Guo, M.; Lim, S.S. Recent advances in recognition of bioactive phytonutrients for specific targets in plant foods. Front. Nutr. 2022, 9, 1018946. [Google Scholar] [CrossRef]
  19. Muscari, I.; Adorisio, S.; Thuy, T.T.; Van Sung, T.; Delfino, D.V. Recent Insights on the Role of Natural Medicines in Immunostimulation. In Natural Medicines; CRC Press: Boca Raton, FL, USA, 2019; pp. 349–360. [Google Scholar]
  20. Heinrich, M.; Barnes, J.; Prieto-Garcia, J.; Gibbons, S.; Williamson, E.M. Natural product chemistry. In Fundamentals of Pharmacognosy and Phytotherapy; Elsevier Health Sciences: Amsterdam, The Netherlands, 2017. [Google Scholar]
  21. Panche, A.N.; Diwan, A.D.; Chandra, S.R. Flavonoids: An overview. J. Nutr. Sci. 2016, 5, e47. [Google Scholar] [CrossRef] [Green Version]
  22. Kumar, S.; Pandey, A.K. Chemistry and biological activities of flavonoids: An overview. Sci. World J. 2013, 2013, 162750. [Google Scholar] [CrossRef] [Green Version]
  23. Ranjan, R.; Logette, E.; Marani, M.; Herzog, M.; Tâche, V.; Scantamburlo, E.; Buchillier, V.; Markram, H. A kinetic map of the homomeric voltage-gated potassium channel (Kv) family. Front. Cell Neurosci. 2019, 13, 358. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. González, C.; Baez-Nieto, D.; Valencia, I.; Oyarzún, I.; Rojas, P.; Naranjo, D.; Latorre, R. K+ channels: Function-structural overview. Compr. Physiol. 2012, 2, 2087–2149. [Google Scholar] [PubMed]
  25. Teisseyre, A.; Michalak, K. Genistein inhibits the activity of Kv1. 3 potassium channels in human T lymphocytes. J. Membr. Biol. 2005, 205, 71–79. [Google Scholar] [CrossRef]
  26. Teisseyre, A.; Palko-Labuz, A.; Uryga, A.; Michalak, K. The influence of 6-Prenylnaringenin and selected non-prenylated flavonoids on the activity of Kv1. 3 channels in human Jurkat T cells. J. Membr. Biol. 2018, 251, 695–704. [Google Scholar] [CrossRef] [PubMed]
  27. Zhao, N.; Dong, Q.; Fu, X.X.; Du, L.L.; Cheng, X.; Du, Y.M.; Liao, Y.H. Acacetin blocks kv1.3 channels and inhibits human T cell activation. Cell. Physiol. Biochem. 2014, 34, 1359–1372. [Google Scholar] [CrossRef]
  28. Teisseyre, A.; Chmielarz, M.; Uryga, A.; Środa Pomianek, K.; Palko-Łabuz, A. Co-Application of Statin and Flavonoids as an Effective Strategy to Reduce the Activity of Voltage-Gated Potassium Channels Kv1.3 and Induce Apoptosis in Human Leukemic T Cell Line Jurkat. Molecules 2022, 27, 3227. [Google Scholar] [CrossRef]
  29. Teisseyre, A.; Michalak, K. Inhibition of the activity of human lymphocyte Kv1.3 potassium channels by resveratrol. J. Membr. Biol. 2006, 214, 123–129. [Google Scholar] [CrossRef] [PubMed]
  30. Teisseyre, A.; Duarte, N.; Ferreira, M.J.U.; Michalak, K. Influence of the multidrug transporter inhibitors on the activity of Kv1.3 voltage-gated potassium channels. Acta Physiol. Pol. 2009, 60, 69. [Google Scholar]
  31. Gąsiorowska, J.; Teisseyre, A.; Uryga, A.; Michalak, K. Inhibition of Kv1.3 channels in human Jurkat T cells by xanthohumol and isoxanthohumol. J. Membr. Biol. 2015, 248, 705–711. [Google Scholar] [CrossRef] [Green Version]
  32. Gasiorowska, J.; Teisseyre, A.; Uryga, A.; Michalak, K. The influence of 8-prenylnaringenin on the activity of voltage-gated kv1.3 potassium channels in human jurkat t cells. Cell. Mol. Biol. Lett. 2012, 17, 559–570. [Google Scholar] [CrossRef]
  33. Phan, H.T.L.; Kim, H.J.; Jo, S.; Kim, W.K.; Namkung, W.; Nam, J.H. Anti-Inflammatory Effect of Licochalcone A via Regulation of ORAI1 and K+ Channels in T-Lymphocytes. Int. J. Mol. Sci. 2021, 22, 10847. [Google Scholar] [CrossRef] [PubMed]
  34. Ou, X.; Bin, X.; Wang, L.; Li, M.; Yang, Y.; Fan, X.; Zeng, X. Myricetin inhibits Kv1.5 channels in HEK293 cells. Mol. Med. Rep. 2016, 13, 1725–1731. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Wang, H.; Wang, H.f.; Wang, C.; Chen, Y.f.; Ma, R.; Xiang, J.z.; Du, X.l.; Tang, Q. Inhibitory effects of hesperetin on Kv1.5 potassium channels stably expressed in HEK 293 cells and ultra-rapid delayed rectifier K+ current in human atrial myocytes. Eur. J. Pharmacol. 2016, 789, 98–108. [Google Scholar] [CrossRef]
  36. Yang, L.; Ma, J.H.; Zhang, P.H.; Zou, A.R.; Tu, D.N. Quercetin activates human Kv1.5 channels by a residue I502 in the S6 segment. Clin. Exp. Pharmacol. Physiol. 2009, 36, 154–161. [Google Scholar] [CrossRef] [PubMed]
  37. Morales-Cano, D.; Menendez, C.; Moreno, E.; Moral-Sanz, J.; Barreira, B.; Galindo, P.; Pandolfi, R.; Jimenez, R.; Moreno, L.; Cogolludo, A.; et al. The flavonoid quercetin reverses pulmonary hypertension in rats. PLoS ONE 2014, 9, e114492. [Google Scholar] [CrossRef] [Green Version]
  38. Liu, Y.; Xu, X.H.; Liu, Z.; Du, X.L.; Chen, K.H.; Xin, X.; Jin, Z.D.; Shen, J.Z.; Hu, Y.; Li, G.R.; et al. Effects of the natural flavone trimethylapigenin on cardiac potassium currents. Biochem. Pharmacol. 2012, 84, 498–506. [Google Scholar] [CrossRef]
  39. Choi, B.H.; Choi, J.S.; Yoon, S.H.; Rhie, D.J.; Jo, Y.H.; Kim, M.S.; Hahn, S.J. Effects of (−)-epigallocatechin-3-gallate, the main component of green tea, on the cloned rat brain Kv1.5 potassium channels. Biochem. Pharmacol. 2001, 62, 527–535. [Google Scholar] [CrossRef]
  40. Noguchi, C.; Yang, J.; Sakamoto, K.; Maeda, R.; Takahashi, K.; Takasugi, H.; Ono, T.; Murakawa, M.; Kimura, J. Inhibitory effects of isoliquiritigenin and licorice extract on voltage-dependent K+ currents in H9c2 cells. J. Pharmacol. Sci. 2008, 108, 439–445. [Google Scholar] [CrossRef] [Green Version]
  41. Wu, H.J.; Wu, W.; Sun, H.Y.; Qin, G.W.; Wang, H.B.; Wang, P.; Yalamanchili, H.K.; Wang, J.; Tse, H.F.; Lau, C.P.; et al. Acacetin causes a frequency-and use-dependent blockade of hKv1.5 channels by binding to the S6 domain. J. Mol. Cell. Cardiol. 2011, 51, 966–973. [Google Scholar] [CrossRef]
  42. Li, G.R.; Wang, H.B.; Qin, G.W.; Jin, M.W.; Tang, Q.; Sun, H.Y.; Du, X.L.; Deng, X.L.; Zhang, X.H.; Chen, J.B.; et al. Acacetin, a natural flavone, selectively inhibits human atrial repolarization potassium currents and prevents atrial fibrillation in dogs. Circulation 2008, 117, 2449–2457. [Google Scholar] [CrossRef]
  43. Xu, H.; Zhao, M.; Liang, S.; Huang, Q.; Xiao, Y.; Ye, L.; Wang, Q.; He, L.; Ma, L.; Zhang, H.; et al. The effects of puerarin on rat ventricular myocytes and the potential mechanism. Sci. Rep. 2016, 6, 35475. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Kang, J.; Cheng, H.; Ji, J.; Incardona, J.; Rampe, D. In vitro electrocardiographic and cardiac ion channel effects of (−)-epigallocatechin-3-gallate, the main catechin of green tea. J. Pharmacol. Exp. Ther. 2010, 334, 619–626. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  45. Sanson, C.; Boukaiba, R.; Houtmann, S.; Maizières, M.A.; Fouconnier, S.; Partiseti, M.; Bohme, G.A. The grapefruit polyphenol naringenin inhibits multiple cardiac ion channels. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2022, 395, 735–740. [Google Scholar] [CrossRef] [PubMed]
  46. Aréchiga-Figueroa, I.A.; Morán-Zendejas, R.; Delgado-Ramírez, M.; Rodríguez-Menchaca, A.A. Phytochemicals genistein and capsaicin modulate Kv2.1 channel gating. Pharmacol. Rep. 2017, 69, 1145–1153. [Google Scholar] [CrossRef] [PubMed]
  47. Gu, R.R.; Meng, X.H.; Zhang, Y.; Xu, H.Y.; Zhan, L.; Gao, Z.B.; Yang, J.L.; Zheng, Y.M. (−)-Naringenin 4′,7-dimethyl Ether Isolated from Nardostachys jatamansi Relieves Pain through Inhibition of Multiple Channels. Molecules 2022, 27, 1735. [Google Scholar] [PubMed]
  48. Kim, H.J.; Ahn, H.S.; Choi, B.H.; Hahn, S.J. Inhibition of Kv4.3 by genistein via a tyrosine phosphorylation-independent mechanism. Am. J. Physiol. Cell Physiol. 2011, 300, C567–C575. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  49. Na, W.; Ma, B.; Shi, S.; Chen, Y.; Zhang, H.; Zhan, Y.; An, H. Procyanidin B1, a novel and specific inhibitor of Kv10.1 channel, suppresses the evolution of hepatoma. Biochem. Pharmacol. 2020, 178, 114089. [Google Scholar] [CrossRef]
  50. Cahalan, M.D.; Chandy, K.G. The functional network of ion channels in T lymphocytes. Immunol. Rev. 2009, 231, 59–87. [Google Scholar] [CrossRef] [Green Version]
  51. DeCoursey, T.E.; Chandy, K.G.; Gupta, S.; Cahalan, M.D. Voltage-gated K+ channels in human T lymphocytes: A role in mitogenesis? Nature 1984, 307, 465–468. [Google Scholar] [CrossRef]
  52. Panyi, G.; Possani, L.; De La Vega, R.R.; Gaspar, R.; Varga, Z. K+ channel blockers: Novel tools to inhibit T cell activation leading to specific immunosuppression. Curr. Pharm. Des. 2006, 12, 2199–2220. [Google Scholar] [CrossRef]
  53. Teisseyre, A.; Gąsiorowska, J.; Michalak, K. Voltage-Gated Potassium Channels Kv1.3–Potentially New Molecular Target in Cancer Diagnostics and Therapy. Adv. Clin. Exp. Med. Off. Organ Wroc. Med. Univ. 2015, 24, 517–524. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Yu, Z.; Li, W.; Liu, F. Inhibition of proliferation and induction of apoptosis by genistein in colon cancer HT-29 cells. Cancer Lett. 2004, 215, 159–166. [Google Scholar] [CrossRef] [PubMed]
  55. Teisseyre, A.; Uryga, A.; Michalak, K. Statins as inhibitors of voltage-gated potassium channels Kv1.3 in cancer cells. J. Mol. Struct. 2021, 1230, 129905. [Google Scholar] [CrossRef]
  56. Baell, J.B.; Gable, R.W.; Harvey, A.J.; Toovey, N.; Herzog, T.; Hänsel, W.; Wulff, H. Khellinone derivatives as blockers of the voltage-gated potassium channel Kv1.3: Synthesis and immunosuppressive activity. J. Med. Chem. 2004, 47, 2326–2336. [Google Scholar] [CrossRef] [PubMed]
  57. Cianci, J.; Baell, J.B.; Flynn, B.L.; Robert, W.G.; Mould, J.A.; Paul, D.; Harvey, A.J. Synthesis and biological evaluation of chalcones as inhibitors of the voltage-gated potassium channel Kv1.3. Bioorg. Med. Chem. Lett. 2008, 18, 2055–2061. [Google Scholar] [CrossRef] [PubMed]
  58. Tamkun, M.M.; Knoth, K.M.; Walbridge, J.A.; Kroemer, H.; Roden, D.M.; Glover, D.M. Molecular cloning and characterization of two voltage-gated K+ channel cDNAs from human ventricle. FASEB J. 1991, 5, 331–337. [Google Scholar] [CrossRef]
  59. Overturf, K.E.; Russell, S.N.; Carl, A.; Vogalis, F.; Hart, P.; Hume, J.; Sanders, K.; Horowitz, B. Cloning and characterization of a Kv1.5 delayed rectifier K+ channel from vascular and visceral smooth muscles. Am. J. Physiol. Cell Physiol. 1994, 267, C1231–C1238. [Google Scholar] [CrossRef]
  60. Wang, Z.; Fermini, B.; Nattel, S. Sustained depolarization-induced outward current in human atrial myocytes. Evidence for a novel delayed rectifier K+ current similar to Kv1.5 cloned channel currents. Circ. Res. 1993, 73, 1061–1076. [Google Scholar] [CrossRef] [Green Version]
  61. Fedida, D.; Wible, B.; Wang, Z.; Fermini, B.; Faust, F.; Nattel, S.; Brown, A. Identity of a novel delayed rectifier current from human heart with a cloned K+ channel current. Circ. Res. 1993, 73, 210–216. [Google Scholar] [CrossRef] [Green Version]
  62. Wettwer, E.; Hála, O.; Christ, T.; Heubach, J.F.; Dobrev, D.; Knaut, M.; Varró, A.; Ravens, U. Role of I Kur in controlling action potential shape and contractility in the human atrium: Influence of chronic atrial fibrillation. Circulation 2004, 110, 2299–2306. [Google Scholar] [CrossRef] [Green Version]
  63. Christophersen, I.E.; Olesen, M.S.; Liang, B.; Andersen, M.N.; Larsen, A.P.; Nielsen, J.B.; Haunsø, S.; Olesen, S.P.; Tveit, A.; Svendsen, J.H.; et al. Genetic variation in KCNA5: Impact on the atrial-specific potassium current I Kur in patients with lone atrial fibrillation. Eur. Heart J. 2013, 34, 1517–1525. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Brendel, J.; Peukert, S. Blockers of the Kv1.5 channel for the treatment of atrial arrhythmias. Expert Opin. Ther. Patents 2002, 12, 1589–1598. [Google Scholar] [CrossRef]
  65. Tamargo, J.; Caballero, R.; Gómez, R.; Delpón, E. IKur/Kv1.5 channel blockers for the treatment of atrial fibrillation. Expert Opin. Investig. Drugs 2009, 18, 399–416. [Google Scholar] [CrossRef]
  66. Ford, J.W.; Milnes, J.T. New drugs targeting the cardiac ultra-rapid delayed-rectifier current (IKur): Rationale, pharmacology and evidence for potential therapeutic value. J. Cardiovasc. Pharmacol. 2008, 52, 105–120. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  67. Van Wagoner, D.R.; Pond, A.L.; McCarthy, P.M.; Trimmer, J.S.; Nerbonne, J.M. Outward K+ current densities and Kv1.5 expression are reduced in chronic human atrial fibrillation. Circ. Res. 1997, 80, 772–781. [Google Scholar] [CrossRef]
  68. He, Y.; Fang, X.; Shi, J.; Li, X.; Xie, M.; Liu, X. Apigenin attenuates pulmonary hypertension by inducing mitochondria-dependent apoptosis of PASMCs via inhibiting the hypoxia inducible factor 1α–KV1.5 channel pathway. Chem. Biol. Interact. 2020, 317, 108942. [Google Scholar] [CrossRef]
  69. Murakoshi, H.; Trimmer, J.S. Identification of the Kv2.1 K+ channel as a major component of the delayed rectifier K+ current in rat hippocampal neurons. J. Neurosci. 1999, 19, 1728–1735. [Google Scholar] [CrossRef] [Green Version]
  70. Malin, S.A.; Nerbonne, J.M. Delayed rectifier K+ currents, IK, are encoded by Kv2 α-subunits and regulate tonic firing in mammalian sympathetic neurons. J. Neurosci. 2002, 22, 10094–10105. [Google Scholar] [CrossRef] [Green Version]
  71. Trimmer, J.S. Immunological identification and characterization of a delayed rectifier K+ channel polypeptide in rat brain. Proc. Natl. Acad. Sci. USA 1991, 88, 10764–10768. [Google Scholar] [CrossRef] [Green Version]
  72. Misonou, H.; Mohapatra, D.P.; Trimmer, J.S. Kv2.1: A voltage-gated k+ channel critical to dynamic control of neuronal excitability. Neurotoxicology 2005, 26, 743–752. [Google Scholar] [CrossRef]
  73. Roe, M.W.; Worley, J.F.; Mittal, A.A.; Kuznetsov, A.; DasGupta, S.; Mertz, R.J.; Witherspoon, S.M.; Blair, N.; Lancaster, M.E.; McIntyre, M.S.; et al. Expression and function of pancreatic β-cell delayed rectifier K+ channels: Role in stimulus-secretion coupling. J. Biol. Chem. 1996, 271, 32241–32246. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Baldwin, T.J.; Tsaur, M.L.; Lopez, G.A.; Jan, Y.N.; Jan, L.Y. Characterization of a mammalian cDNA for an inactivating voltage-sensitive K+ channel. Neuron 1991, 7, 471–483. [Google Scholar] [CrossRef] [PubMed]
  75. Jerng, H.H.; Pfaffinger, P.J.; Covarrubias, M. Molecular physiology and modulation of somatodendritic A-type potassium channels. Mol. Cell. Neurosci. 2004, 27, 343–369. [Google Scholar] [CrossRef] [PubMed]
  76. Pak, M.D.; Baker, K.; Covarrubias, M.; Butler, A.; Ratcliffe, A.; Salkoff, L. mShal, a subfamily of A-type K+ channel cloned from mammalian brain. Proc. Natl. Acad. Sci. USA 1991, 88, 4386–4390. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  77. Guo, Y.; Zhang, C.; Ye, T.; Chen, X.; Liu, X.; Chen, X.; Sun, Y.; Qu, C.; Liang, J.; Shi, S.; et al. Pinocembrin ameliorates arrhythmias in rats with chronic ischaemic heart failure. Ann. Med. 2021, 53, 830–840. [Google Scholar] [CrossRef] [PubMed]
  78. Chen, X.; Wan, W.; Ran, Q.; Ye, T.; Sun, Y.; Liu, Z.; Liu, X.; Shi, S.; Qu, C.; Zhang, C.; et al. Pinocembrin mediates antiarrhythmic effects in rats with isoproterenol-induced cardiac remodeling. Eur. J. Pharmacol. 2022, 920, 174799. [Google Scholar] [CrossRef] [PubMed]
  79. Haverkamp, W.; Breithardt, G.; Camm, A.J.; Janse, M.J.; Rosen, M.R.; Antzelevitch, C.; Escande, D.; Franz, M.; Malik, M.; Moss, A.; et al. The potential for QT prolongation and pro-arrhythmia by non-anti-arrhythmic drugs: Clinical and regulatory implications: Report on a Policy Conference of the European Society of Cardiology. Cardiovasc. Res. 2000, 47, 219–233. [Google Scholar] [CrossRef]
  80. Taglialatela, M.; Castaldo, P.; Pannaccione, A.; Giorgio, G.; Annunziato, L. Human ether-a-gogo related gene (HERG) K channels as pharmacological targets: Present and future implications. Biochem. Pharmacol. 1998, 55, 1741–1746. [Google Scholar] [CrossRef]
  81. Saponara, S.; Fusi, F.; Iovinelli, D.; Ahmed, A.; Trezza, A.; Spiga, O.; Sgaragli, G.; Valoti, M. Flavonoids and hERG channels: Friends or foes? Eur. J. Pharmacol. 2021, 899, 174030. [Google Scholar] [CrossRef]
  82. Zitron, E.; Scholz, E.; Owen, R.W.; Luück, S.; Kiesecker, C.; Thomas, D.; Kathoöfer, S.; Niroomand, F.; Kiehn, J.; Kreye, V.A.; et al. QTc prolongation by grapefruit juice and its potential pharmacological basis: HERG channel blockade by flavonoids. Circulation 2005, 111, 835–838. [Google Scholar] [CrossRef]
  83. Scholz, E.P.; Zitron, E.; Kiesecker, C.; Lück, S.; Thomas, D.; Kathöfer, S.; Kreye, V.A.; Katus, H.A.; Kiehn, J.; Schoels, W.; et al. Inhibition of cardiac HERG channels by grapefruit flavonoid naringenin: Implications for the influence of dietary compounds on cardiac repolarisation. Naunyn-Schmiedeberg’s Arch. Pharmacol. 2005, 371, 516–525. [Google Scholar] [CrossRef] [PubMed]
  84. Lin, C.; Ke, X.; Ranade, V.; Somberg, J. The additive effects of the active component of grapefruit juice (naringenin) and antiarrhythmic drugs on HERG inhibition. Cardiology 2008, 110, 145–152. [Google Scholar] [CrossRef] [PubMed]
  85. Zhang, D.Y.; Wang, Y.; Lau, C.P.; Tse, H.F.; Li, G.R. Both EGFR kinase and Src-related tyrosine kinases regulate human ether-a-go-go-related gene potassium channels. Cell. Signal. 2008, 20, 1815–1821. [Google Scholar] [CrossRef] [PubMed]
  86. Du, K.; De Mieri, M.; Saxena, P.; Phungula, K.V.; Wilhelm, A.; Hrubaru, M.M.; van Rensburg, E.; Zietsman, P.C.; Hering, S.; van der Westhuizen, J.H.; et al. HPLC-Based activity profiling for Herg channel inhibitors in the South African medicinal plant Galenia africana. Planta Med. 2015, 81, 1154–1162. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Kelemen, K.; Kiesecker, C.; Zitron, E.; Bauer, A.; Scholz, E.; Bloehs, R.; Thomas, D.; Greten, J.; Remppis, A.; Schoels, W.; et al. Green tea flavonoid epigallocatechin-3-gallate (EGCG) inhibits cardiac hERG potassium channels. Biochem. Biophys. Res. Commun. 2007, 364, 429–435. [Google Scholar] [CrossRef] [PubMed]
  88. Sun, X.; Xu, B.; Xue, Y.; Li, H.; Zhang, H.; Zhang, Y.; Kang, L.; Zhang, X.; Zhang, J.; Jia, Z.; et al. Characterization and structure-activity relationship of natural flavonoids as hERG K+ channel modulators. Int. Immunopharmacol. 2017, 45, 187–193. [Google Scholar] [CrossRef] [PubMed]
  89. Scholz, E.P.; Zitron, E.; Kiesecker, C.; Thomas, D.; Kathöfer, S.; Kreuzer, J.; Bauer, A.; Katus, H.A.; Remppis, A.; Karle, C.A.; et al. Orange flavonoid hesperetin modulates cardiac hERG potassium channel via binding to amino acid F656. Nutr. Metab. Cardiovasc. Dis. 2007, 17, 666–675. [Google Scholar] [CrossRef] [PubMed]
  90. Sun, X.H.; Ding, J.P.; Li, H.; Pan, N.; Gan, L.; Yang, X.L.; Xu, H.B. Activation of large-conductance calcium-activated potassium channels by puerarin: The underlying mechanism of puerarin-mediated vasodilation. J. Pharmacol. Exp. Ther. 2007, 323, 391–397. [Google Scholar] [CrossRef] [Green Version]
  91. Sweeney, O.; Wang, T.; Ellory, C.; Wilkins, R.; Ma, Y. The effects of liquiritigenin on the activity of the hERG potassium channel. Br. J. Pharmacol. 2019, 176, 3067–3068. [Google Scholar]
  92. Yun, J.; Bae, H.; Choi, S.E.; Kim, J.H.; Choi, Y.W.; Lim, I.; Lee, C.S.; Lee, M.W.; Ko, J.H.; Seo, S.J.; et al. Taxifolin Glycoside Blocks Human ether-a-go-go Related Gene K+ Channels. Korean J. Physiol. Pharmacol. Off. J. Korean Physiol. Soc. Korean Soc. Pharmacol. 2013, 17, 37. [Google Scholar] [CrossRef] [Green Version]
  93. Sanguinetti, M.C.; Jurkiewicz, N.K. Two components of cardiac delayed rectifier K+ current. Differential sensitivity to block by class III antiarrhythmic agents. J. Gen. Physiol. 1990, 96, 195–215. [Google Scholar] [CrossRef] [PubMed]
  94. Liu, Y.; Zhang, L.; Dong, L.; Song, Q.; Guo, P.; Wang, Y.; Chen, Z.; Zhang, M. Hesperetin improves diabetic coronary arterial vasomotor responsiveness by upregulating myocyte voltage-gated K+ channels. Exp. Ther. Med. 2020, 20, 486–494. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  95. Wang, S.; Zhang, S.; Wang, S.; Gao, P.; Dai, L. A comprehensive review on Pueraria: Insights on its chemistry and medicinal value. Biomed. Pharmacother. 2020, 131, 110734. [Google Scholar] [CrossRef] [PubMed]
  96. Ouadid-Ahidouch, H.; Ahidouch, A.; Pardo, L.A. Kv10.1 K+ channel: From physiology to cancer. PflÜGers-Arch. Eur. J. Physiol. 2016, 468, 751–762. [Google Scholar] [CrossRef]
  97. Marty, A. Ca-dependent K channels with large unitary conductance in chromaffin cell membranes. Nature 1981, 291, 497–500. [Google Scholar] [CrossRef]
  98. Cui, J.; Yang, H.; Lee, U.S. Molecular mechanisms of BK channel activation. Cell. Mol. Life Sci. 2009, 66, 852–875. [Google Scholar] [CrossRef] [Green Version]
  99. Kulawiak, B.; Kudin, A.P.; Szewczyk, A.; Kunz, W.S. BK channel openers inhibit ROS production of isolated rat brain mitochondria. Exp. Neurol. 2008, 212, 543–547. [Google Scholar] [CrossRef]
  100. Szabo, I.; Zoratti, M. Mitochondrial channels: Ion fluxes and more. Physiol. Rev. 2014, 94, 519–608. [Google Scholar] [CrossRef]
  101. Krabbendam, I.E.; Honrath, B.; Culmsee, C.; Dolga, A.M. Mitochondrial Ca2+-activated K+ channels and their role in cell life and death pathways. Cell Calcium 2018, 69, 101–111. [Google Scholar] [CrossRef]
  102. Saponara, S.; Testai, L.; Iozzi, D.; Martinotti, E.; Martelli, A.; Chericoni, S.; Sgaragli, G.; Fusi, F.; Calderone, V. (+/−)-Naringenin as large conductance Ca2+-activated K+ (BKCa) channel opener in vascular smooth muscle cells. Br. J. Pharmacol. 2006, 149, 1013–1021. [Google Scholar] [CrossRef] [Green Version]
  103. Hsu, H.T.; Tseng, Y.T.; Lo, Y.C.; Wu, S.N. Ability of naringenin, a bioflavonoid, to activate M-type potassium current in motor neuron-like cells and to increase BKCa-channel activity in HEK293T cells transfected with α-hSlo subunit. BMC Neurosci. 2014, 15, 135. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  104. Yang, Z.; Pan, A.; Zuo, W.; Guo, J.; Zhou, W. Relaxant effect of flavonoid naringenin on contractile activity of rat colonic smooth muscle. J. Ethnopharmacol. 2014, 155, 1177–1183. [Google Scholar] [CrossRef] [PubMed]
  105. Shi, R.; Xu, J.W.; Xiao, Z.T.; Chen, R.F.; Zhang, Y.L.; Lin, J.B.; Cheng, K.L.; Wei, G.Y.; Li, P.B.; Zhou, W.L.; et al. Naringin and naringenin relax rat tracheal smooth by regulating BKCa activation. J. Med. Food 2019, 22, 963–970. [Google Scholar] [CrossRef] [PubMed]
  106. Testai, L.; Martelli, A.; Marino, A.; D’antongiovanni, V.; Ciregia, F.; Giusti, L.; Lucacchini, A.; Chericoni, S.; Breschi, M.C.; Calderone, V. The activation of mitochondrial BK potassium channels contributes to the protective effects of naringenin against myocardial ischemia/reperfusion injury. Biochem. Pharmacol. 2013, 85, 1634–1643. [Google Scholar] [CrossRef] [PubMed]
  107. Testai, L.; Da Pozzo, E.; Piano, I.; Pistelli, L.; Gargini, C.; Breschi, M.C.; Braca, A.; Martini, C.; Martelli, A.; Calderone, V. The citrus flavanone naringenin produces cardioprotective effects in hearts from 1 year old rat, through activation of mitoBK channels. Front. Pharmacol. 2017, 8, 71. [Google Scholar] [CrossRef] [Green Version]
  108. Kampa, R.P.; Kicinska, A.; Jarmuszkiewicz, W.; Pasikowska-Piwko, M.; Dolegowska, B.; Debowska, R.; Szewczyk, A.; Bednarczyk, P. Naringenin as an opener of mitochondrial potassium channels in dermal fibroblasts. Exp. Dermatol. 2019, 28, 543–550. [Google Scholar] [CrossRef]
  109. Kicinska, A.; Kampa, R.P.; Daniluk, J.; Sek, A.; Jarmuszkiewicz, W.; Szewczyk, A.; Bednarczyk, P. Regulation of the mitochondrial BKCa channel by the citrus flavonoid naringenin as a potential means of preventing cell damage. Molecules 2020, 25, 3010. [Google Scholar] [CrossRef]
  110. Da Pozzo, E.; Costa, B.; Cavallini, C.; Testai, L.; Martelli, A.; Calderone, V.; Martini, C. The citrus flavanone naringenin protects myocardial cells against age-associated damage. Oxidative Med. Cell. Longev. 2017, 2017, 9536148. [Google Scholar] [CrossRef] [Green Version]
  111. Richter-Laskowska, M.; Trybek, P.; Bednarczyk, P.; Wawrzkiewicz-Jałowiecka, A. To what extent naringenin binding and membrane depolarization shape mitoBK channel gating—A machine learning approach. PLoS Comput. Biol. 2022, 18, e1010315. [Google Scholar] [CrossRef]
  112. Côrtes, S.F.; Rezende, B.A.; Corriu, C.; Medeiros, I.A.; Teixeira, M.M.; Lopes, M.J.; Lemos, V.S. Pharmacological evidence for the activation of potassium channels as the mechanism involved in the hypotensive and vasorelaxant effect of dioclein in rat small resistance arteries. Br. J. Pharmacol. 2001, 133, 849. [Google Scholar] [CrossRef]
  113. Dimpfel, W. Different anticonvulsive effects of hesperidin and its aglycone hesperetin on electrical activity in the rat hippocampus in-vitro. J. Pharm. Pharmacol. 2006, 58, 375–379. [Google Scholar] [CrossRef] [PubMed]
  114. Kim, Y.; Kim, W.J.; Cha, E.J. Quercetin-induced growth inhibition in human bladder cancer cells is associated with an increase in Ca2+-activated K+ channels. Korean J. Physiol. Pharmacol. 2011, 15, 279–283. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  115. Melnyk, M.I.; Dryn, D.O.; Al Kury, L.T.; Zholos, A.V.; Soloviev, A.I. Liposomal quercetin potentiates maxi-K channel openings in smooth muscles and restores its activity after oxidative stress. J. Liposome Res. 2019, 29, 94–101. [Google Scholar] [CrossRef] [PubMed]
  116. Zhang, Y.M.; Zhang, Z.Y.; Wang, R.X. Protective mechanisms of quercetin against myocardial ischemia reperfusion injury. Front. Physiol. 2020, 11, 956. [Google Scholar] [CrossRef] [PubMed]
  117. Kampa, R.P.; Sek, A.; Szewczyk, A.; Bednarczyk, P. Cytoprotective effects of the flavonoid quercetin by activating mitochondrial BKCa channels in endothelial cells. Biomed. Pharmacother. 2021, 142, 112039. [Google Scholar] [CrossRef]
  118. Kampa, R.P.; Gliździńska, A.; Szewczyk, A.; Bednarczyk, P.; Filipek, S. Flavonoid quercetin abolish paxilline inhibition of the mitochondrial BKCa channel. Mitochondrion 2022, 65, 23–32. [Google Scholar] [CrossRef]
  119. Li, Y.; Starrett, J.E.; Meanwell, N.A.; Johnson, G.; Harte, W.E.; Dworetzky, S.I.; Boissard, C.G.; Gribkoff, V.K. The discovery of novel openers of Ca2+-dependent large-conductance potassium channels: Pharmacophore search and physiological evaluation of flavonoids. Bioorg. Med. Chem. Lett. 1997, 7, 759–762. [Google Scholar] [CrossRef]
  120. Mahobiya, A.; Singh, T.U.; Rungsung, S.; Kumar, T.; Chandrasekaran, G.; Parida, S.; Kumar, D. Kaempferol-induces vasorelaxation via endothelium-independent pathways in rat isolated pulmonary artery. Pharmacol. Rep. 2018, 70, 863–874. [Google Scholar] [CrossRef]
  121. Xu, Y.; Leung, G.; Wong, P.; Vanhoutte, P.; Man, R. Kaempferol stimulates large conductance Ca2+-activated K+ (BKCa) channels in human umbilical vein endothelial cells via a cAMP/PKA-dependent pathway. Br. J. Pharmacol. 2008, 154, 1247–1253. [Google Scholar] [CrossRef] [Green Version]
  122. Xu, Y.; Leung, S.; Leung, G.; Man, R. Kaempferol enhances endothelium-dependent relaxation in the porcine coronary artery through activation of large-conductance Ca 2+-activated K+ channels. Br. J. Pharmacol. 2015, 172, 3003–3014. [Google Scholar] [CrossRef] [Green Version]
  123. Kampa, R.P.; Flori, L.; Sęk, A.; Spezzini, J.; Brogi, S.; Szewczyk, A.; Calderone, V.; Bednarczyk, P.; Testai, L. Luteolin-Induced Activation of Mitochondrial BKCa Channels: Undisclosed Mechanism of Cytoprotection. Antioxidants 2022, 11, 1892. [Google Scholar] [CrossRef] [PubMed]
  124. Sun, L.; Gonzalez, L.A.; Horrigan, F.T. Nobiletin Inhibition of BK Channels. Biophys. J. 2019, 116, 104a. [Google Scholar] [CrossRef]
  125. Saadat, S.; Boskabadi, J.; Boskabady, M.H. Contribution of potassium channels, beta2-adrenergic and histamine H1 receptors in the relaxant effect of baicalein on rat tracheal smooth muscle. Iran. J. Basic Med. Sci. 2019, 22, 1347. [Google Scholar] [PubMed]
  126. Lamai, J.; Mahabusarakam, W.; Ratithammatorn, T.; Hiranyachattada, S. Effects of morelloflavone from Garcinia dulcis on vasorelaxation of isolated rat thoracic aorta. J. Physiol. Biomed. Sci. 2013, 26, 13–17. [Google Scholar]
  127. Bai, B.; Lu, N.; Zhang, W.; Lin, J.; Zhao, T.; Zhou, S.; Khasanova, E.; Zhang, L. Inhibitory Effects of Genistein on Vascular Smooth Muscle Cell Proliferation Induced by Ox-LDL: Role of BKCa Channels. Anal. Cell. Pathol. 2020, 2020, 8895449. [Google Scholar] [CrossRef]
  128. Sun, L.; Zhao, T.; Ju, T.; Wang, X.; Li, X.; Wang, L.; Zhang, L.; Yu, G. A combination of intravenous genistein plus Mg2+ enhances antihypertensive effects in SHR by endothelial protection and BKCa channel inhibition. Am. J. Hypertens. 2015, 28, 1114–1120. [Google Scholar] [CrossRef] [Green Version]
  129. Zhou, R.; Liu, L.; Hu, D. Involvement of BKCa α subunit tyrosine phosphorylation in vascular hyporesponsiveness of superior mesenteric artery following hemorrhagic shock in rats. Cardiovasc. Res. 2005, 68, 327–335. [Google Scholar] [CrossRef]
  130. Wang, Y.; Sun, H.Y.; Liu, Y.G.; Song, Z.; She, G.; Xiao, G.S.; Wang, Y.; Li, G.R.; Deng, X.L. Tyrphostin AG 556 increases the activity of large conductance Ca2+-activated K+ channels by inhibiting epidermal growth factor receptor tyrosine kinase. J. Cell. Mol. Med. 2017, 21, 1826–1834. [Google Scholar] [CrossRef]
  131. Stumpff, F.; Que, Y.; Boxberger, M.; Strauss, O.; Wiederholt, M. Stimulation of maxi-K channels in trabecular meshwork by tyrosine kinase inhibitors. Investig. Ophthalmol. Vis. Sci. 1999, 40, 1404–1417. [Google Scholar]
  132. Zhang, H.T.; Wang, Y.; Deng, X.L.; Dong, M.Q.; Zhao, L.M.; Wang, Y.W. Daidzein relaxes rat cerebral basilar artery via activation of large-conductance Ca2+-activated K+ channels in vascular smooth muscle cells. Eur. J. Pharmacol. 2010, 630, 100–106. [Google Scholar] [CrossRef]
  133. Nevala, R.; Paukku, K.; Korpela, R.; Vapaatalo, H. Calcium-sensitive potassium channel inhibitors antagonize genistein-and daidzein-induced arterial relaxation in vitro. Life Sci. 2001, 69, 1407–1417. [Google Scholar] [CrossRef] [PubMed]
  134. Yang, B.; Gao, Q.; Yao, H.; Xia, Q. Mitochondrial mechanism of cardioprotective effect of puerarin against H2O2-stress in rats. Zhongguo Ying Yong Sheng Xue Zhi = Zhongguo Yingyong Shenglixue Zazhi = Chin. J. Appl. Physiol. 2008, 24, 399–404. [Google Scholar]
  135. Yao, H.; Gao, Q.; Xia, Q. The role of mitochondrial K+ channels in the cardioprotection of puerarin against hypoxia/reoxygenation injury in rats. Zhongguo Ying Yong Sheng Xue Zhi = Zhongguo Yingyong Shenglixue Zazhi = Chin. J. Appl. Physiol. 2010, 26, 459–462. [Google Scholar]
  136. Koh, D.S.; Reid, G.; Vogel, W. Effect of the flavoid phloretin on Ca2+-activated K+ channels in myelinated nerve fibres of Xenopus laevis. Neurosci. Lett. 1994, 165, 167–170. [Google Scholar] [CrossRef] [PubMed]
  137. Gonzalez, L.A.; Ma, Z.; Horrigan, F.T. The BK Channel Opener Phloretin Influences Voltage-and Calcium-Dependent Gating. Biophys. J. 2012, 102, 683a–684a. [Google Scholar] [CrossRef] [Green Version]
  138. Gonzalez, L.A.; Ma, Z.; Horrigan, F.T. Potential Sites of Action for Phloretin on BK Channels. Biophys. J. 2013, 104, 463a. [Google Scholar] [CrossRef] [Green Version]
  139. Marques, A.A.M.; da Silva, C.H.F.; de Souza, P.; de Almeida, C.L.; Cechinel-Filho, V.; Lourenço, E.L.; Junior, A.G. Nitric oxide and Ca2+-activated high-conductance K+ channels mediate nothofagin-induced endothelium-dependent vasodilation in the perfused rat kidney. Chem. Biol. Interact. 2020, 327, 109182. [Google Scholar] [CrossRef]
  140. Teodoro, J.S.; Duarte, F.V.; Rolo, A.P.; Palmeira, C.M. Chapter 28—Mitochondria as a Target for Safety and Toxicity Evaluation of Nutraceuticals. In Nutraceuticals; Gupta, R.C., Ed.; Academic Press: Boston, MA, USA, 2016; pp. 387–400. [Google Scholar] [CrossRef]
  141. Ma, Y.G.; Liang, L.; Zhang, Y.B.; Wang, B.F.; Bai, Y.G.; Dai, Z.J.; Xie, M.J.; Wang, Z.W. Berberine reduced blood pressure and improved vasodilation in diabetic rats. J. Mol. Endocrinol. 2017, 59, 191–204. [Google Scholar] [CrossRef] [Green Version]
  142. Wu, J.; Nakashima, S.; Shigyo, M.; Yamasaki, M.; Ikuno, S.; Morikawa, A.; Takegami, S.; Nakamura, S.; Konishi, A.; Kitade, T.; et al. Antihypertensive constituents in Sanoshashinto. J. Nat. Med. 2020, 74, 421–433. [Google Scholar] [CrossRef]
  143. Cordeiro, B.; Shinn, C.; Sellke, F.W.; Clements, R.T. Rottlerin-induced BKCa channel activation impairs specific contractile responses and promotes vasodilation. Ann. Thorac. Surg. 2015, 99, 626–634. [Google Scholar] [CrossRef]
  144. Goldklang, M.P.; Perez-Zoghbi, J.F.; Trischler, J.; Nkyimbeng, T.; Zakharov, S.I.; Shiomi, T.; Zelonina, T.; Marks, A.R.; D’Armiento, J.M.; Marx, S.O. Treatment of experimental asthma using a single small molecule with anti-inflammatory and BK channel-activating properties. FASEB J. 2013, 27, 4975. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  145. Yang, L.; Han, B.; Zhang, M.; Wang, Y.H.; Tao, K.; Zhu, M.X.; He, K.; Zhang, Z.G.; Hou, S. Activation of BK channels prevents hepatic stellate cell activation and liver fibrosis through the suppression of TGFβ1/SMAD3 and JAK/STAT3 profibrotic signaling pathways. Front. Pharmacol. 2020, 11, 165. [Google Scholar] [CrossRef] [PubMed]
  146. Nishida, S.; Satoh, H. Possible involvement of Ca2+ activated K+ channels, SK channel, in the quercetin-induced vasodilatation. Korean J. Physiol. Pharmacol. 2009, 13, 361–365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  147. Maaliki, D.; Shaito, A.A.; Pintus, G.; El-Yazbi, A.; Eid, A.H. Flavonoids in hypertension: A brief review of the underlying mechanisms. Curr. Opin. Pharmacol. 2019, 45, 57–65. [Google Scholar] [CrossRef]
  148. Calderone, V.; Chericoni, S.; Martinelli, C.; Testai, L.; Nardi, A.; Morelli, I.; Breschi, M.C.; Martinotti, E. Vasorelaxing effects of flavonoids: Investigation on the possible involvement of potassium channels. Naunyn-Schmiedeberg’S Arch. Pharmacol. 2004, 370, 290–298. [Google Scholar] [CrossRef]
  149. Lopes, K.S.; Marques, A.A.M.; Moreno, K.G.T.; Lorençone, B.R.; Leite, P.R.T.; da Silva, G.P.; Dos Santos, A.C.; Souza, R.I.C.; Gasparotto, F.M.; Cassemiro, N.S.; et al. Small conductance calcium-activated potassium channels and nitric oxide/cGMP pathway mediate cardioprotective effects of Croton urucurana Baill. In hypertensive rats. J. Ethnopharmacol. 2022, 293, 115255. [Google Scholar] [CrossRef]
  150. Oliani, J.; Ferreira, M.J.P.; Salatino, A.; Salatino, M.L.F. Leaf flavonoids from Croton urucurana and C. floribundus (Euphorbiaceae). Biochem. Syst. Ecol. 2021, 94, 104217. [Google Scholar] [CrossRef]
  151. da Rocha Lapa, F.; Soares, K.C.; Rattmann, Y.D.; Crestani, S.; Missau, F.C.; Pizzolatti, M.G.; Marques, M.C.A.; Rieck, L.; Santos, A.R.S. Vasorelaxant and hypotensive effects of the extract and the isolated flavonoid rutin obtained from Polygala paniculata L. J. Pharm. Pharmacol. 2011, 63, 875–881. [Google Scholar] [CrossRef]
  152. Chen, K.H.; Liu, H.; Sun, H.Y.; Jin, M.W.; Xiao, G.S.; Wang, Y.; Li, G.R. The natural flavone acacetin blocks small conductance Ca2+-activated K+ channels stably expressed in HEK 293 cells. Front. Pharmacol. 2017, 8, 716. [Google Scholar] [CrossRef] [Green Version]
  153. Hsueh, C.H.; Chang, P.C.; Hsieh, Y.C.; Reher, T.; Chen, P.S.; Lin, S.F. Proarrhythmic effect of blocking the small conductance calcium activated potassium channel in isolated canine left atrium. Heart Rhythm. 2013, 10, 891–898. [Google Scholar] [CrossRef] [Green Version]
  154. Tirloni, C.A.S.; Palozi, R.A.C.; Schaedler, M.I.; Guarnier, L.P.; Silva, A.O.; Marques, M.A.; Gasparotto, F.M.; Lourenço, E.L.B.; de Souza, L.M.; Junior, A.G. Influence of Luehea divaricata Mart. extracts on peripheral vascular resistance and the role of nitric oxide and both Ca+2-sensitive and Kir6. 1 ATP-sensitive K+ channels in the vasodilatory effects of isovitexin on isolated perfused mesenteric beds. Phytomedicine 2019, 56, 74–82. [Google Scholar] [CrossRef] [PubMed]
  155. Reimann, F.; Ashcroft, F.M. Inwardly rectifying potassium channels. Curr. Opin. Cell Biol. 1999, 11, 503–508. [Google Scholar] [CrossRef] [PubMed]
  156. Hibino, H.; Inanobe, A.; Furutani, K.; Murakami, S.; Findlay, I.; Kurachi, Y. Inwardly rectifying potassium channels: Their structure, function, and physiological roles. Physiol. Rev. 2010, 90, 291–366. [Google Scholar] [CrossRef] [Green Version]
  157. Jiao, Y.; Fan, Y.F.; Wang, Y.L.; Zhang, J.Y.; Chen, S.; Chen, Z.W. Protective effect and mechanism of total flavones from Rhododendron simsii planch flower on cultured rat cardiomyocytes with anoxia and reoxygenation. Evid. Based Complement. Altern. Med. 2015, 2015, 863531. [Google Scholar] [CrossRef] [Green Version]
  158. Li, W.; Dong, M.; Guo, P.; Liu, Y.; Jing, Y.; Chen, R.; Zhang, M. Luteolin-induced coronary arterial relaxation involves activation of the myocyte voltage-gated K+ channels and inward rectifier K+ channels. Life Sci. 2019, 221, 233–240. [Google Scholar] [CrossRef] [PubMed]
  159. Trezza, A.; Cicaloni, V.; Porciatti, P.; Langella, A.; Fusi, F.; Saponara, S.; Spiga, O. From in silico to in vitro: A trip to reveal flavonoid binding on the Rattus norvegicus Kir6.1 ATP-sensitive inward rectifier potassium channel. PeerJ 2018, 6, e4680. [Google Scholar] [CrossRef] [Green Version]
  160. Rameshrad, M.; Omidkhoda, S.F.; Razavi, B.M.; Hosseinzadeh, H. Evaluating the possible role of mitochondrial ATP-sensitive potassium channels in the cardioprotective effects of morin in the isolated rat heart. Life Sci. 2021, 264, 118659. [Google Scholar] [CrossRef]
  161. Khan, A.u.; Gilani, A.H. Selective bronchodilatory effect of Rooibos tea (Aspalathus linearis) and its flavonoid, chrysoeriol. Eur. J. Nutr. 2006, 45, 463–469. [Google Scholar] [CrossRef]
  162. Jin, J.Y.; Park, S.H.; Bae, J.H.; Cho, H.C.; Lim, J.G.; Park, W.S.; Han, J.; Lee, J.H.; Song, D.K. Uncoupling by (−)-epigallocatechin-3-gallate of ATP-sensitive potassium channels from phosphatidylinositol polyphosphates and ATP. Pharmacol. Res. 2007, 56, 237–247. [Google Scholar] [CrossRef]
  163. Pinho-Ribeiro, F.A.; Zarpelon, A.C.; Fattori, V.; Manchope, M.F.; Mizokami, S.S.; Casagrande, R.; Verri, W.A., Jr. Naringenin reduces inflammatory pain in mice. Neuropharmacology 2016, 105, 508–519. [Google Scholar] [CrossRef]
  164. Manchope, M.F.; Calixto-Campos, C.; Coelho-Silva, L.; Zarpelon, A.C.; Pinho-Ribeiro, F.A.; Georgetti, S.R.; Baracat, M.M.; Casagrande, R.; Verri, W.A., Jr. Naringenin inhibits superoxide anion-induced inflammatory pain: Role of oxidative stress, cytokines, Nrf-2 and the NO- cGMP- PKG- KATPChannel signaling pathway. PLoS ONE 2016, 11, e0153015. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  165. Moradi, N.; Fakhri, S.; Farzaei, M.H.; Abbaszadeh, F. The anti-nociceptive activity of naringenin passes through L-arginine/NO/cGMP/KATP channel pathway and opioid receptors. Behav. Pharmacol. 2021, 32, 590–598. [Google Scholar] [CrossRef] [PubMed]
  166. Meng, L.M.; Ma, H.J.; Guo, H.; Kong, Q.Q.; Zhang, Y. The cardioprotective effect of naringenin against ischemia–Reperfusion injury through activation of ATP-sensitive potassium channel in rat. Can. J. Physiol. Pharmacol. 2016, 94, 973–978. [Google Scholar] [CrossRef] [PubMed]
  167. Hua, Z.; Wang, X. Inhibitory effect of berberine on potassium channels in guinea pig ventricular myocytes. Yao Xue Xue Bao = Acta Pharm. Sin. 1994, 29, 576–580. [Google Scholar]
  168. Wang, Y.X.; Zheng, Y.M.; Zhou, X.B. Inhibitory effects of berberine on ATP-sensitive K+ channels in cardiac myocytes. Eur. J. Pharmacol. 1996, 316, 307–315. [Google Scholar] [CrossRef]
  169. Suantawee, T.; Elazab, S.T.; Hsu, W.H.; Yao, S.; Cheng, H.; Adisakwattana, S. Cyanidin stimulates insulin secretion and pancreatic β-cell gene expression through activation of L-type voltage-dependent Ca2+ channels. Nutrients 2017, 9, 814. [Google Scholar] [CrossRef]
  170. Ribeiro, A.R.S.; do Nascimento Valença, J.D.; da Silva Santos, J.; Boeing, T.; da Silva, L.M.; de Andrade, S.F.; Albuquerque-Júnior, R.L.; Thomazzi, S.M. The effects of baicalein on gastric mucosal ulcerations in mice: Protective pathways and anti-secretory mechanisms. Chem. Biol. Interact. 2016, 260, 33–41. [Google Scholar] [CrossRef]
  171. Zhao, Z.; Liu, B.; Zhang, G.; Jia, Z.; Jia, Q.; Geng, X.; Zhang, H. Molecular basis for genistein-induced inhibition of Kir2.3 currents. PflÜGers-Arch. Eur. J. Physiol. 2008, 456, 413–423. [Google Scholar] [CrossRef]
  172. Ko, E.A.; Park, W.S.; Son, Y.K.; Kim, D.H.; Kim, N.; Kim, H.K.; Choi, T.H.; Jung, I.D.; Park, Y.M.; Han, J. The effect of tyrosine kinase inhibitor genistein on voltage-dependent K+ channels in rabbit coronary arterial smooth muscle cells. Vasc. Pharmacol. 2009, 50, 51–56. [Google Scholar] [CrossRef]
  173. Okamoto, F.; Okabe, K.; Kajiya, H. Genistein, a soybean isoflavone, inhibits inward rectifier K+ channels in rat osteoclasts. Jpn. J. Physiol. 2001, 51, 501–509. [Google Scholar] [CrossRef] [Green Version]
  174. Ogata, R.; Kitamura, K.; Ito, Y.; Nakano, H. Inhibitory effects of genistein on ATP-sensitive K+ channels in rabbit portal vein smooth muscle. Br. J. Pharmacol. 1997, 122, 1395–1404. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  175. Kampa, R.P.; Sęk, A.; Bednarczyk, P.; Szewczyk, A.; Calderone, V.; Testai, L. Flavonoids as new regulators of mitochondrial potassium channels: Ccontribution to cardioprotection. J. Pharm. Pharmacol. 2022, rgac093. [Google Scholar] [CrossRef] [PubMed]
  176. Song, D.K.; Jang, Y.; Kim, J.H.; Chun, K.J.; Lee, D.; Xu, Z. Polyphenol (−)-epigallocatechin gallate during ischemia limits infarct size via mitochondrial KATP channel activation in isolated rat hearts. J. Korean Med. Sci. 2010, 25, 380–386. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Ma, H.; Huang, X.; Li, Q.; Guan, Y.; Yuan, F.; Zhang, Y. ATP-dependent potassium channels and mitochondrial permeability transition pores play roles in the cardioprotection of theaflavin in young rat. J. Physiol. Sci. 2011, 61, 337–342. [Google Scholar] [CrossRef] [PubMed]
  178. Hu, Y.; Li, L.; Yin, W.; Shen, L.; You, B.; Gao, H. Protective effect of proanthocyanidins on anoxia-reoxygenation injury of myocardial cells mediated by the PI3K/Akt/GSK-3β pathway and mitochondrial ATP-sensitive potassium channel. Mol. Med. Rep. 2014, 10, 2051–2058. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  179. Couvreur, N.; Tissier, R.; Pons, S.; Chenoune, M.; Waintraub, X.; Berdeaux, A.; Ghaleh, B. The Ceiling Effect of Pharmacological Postconditioning with the Phytoestrogen Genistein Is Reversed by the GSK3β Inhibitor SB 216763 [3-(2, 4-Dichlorophenyl)-4 (1-methyl-1H-indol-3-yl)-1H-pyrrole-2, 5-dione] through Mitochondrial ATP-Dependent Potassium Channel Opening. J. Pharmacol. Exp. Ther. 2009, 329, 1134–1141. [Google Scholar]
  180. Tu, I.H.; Yen, H.T.D.; Cheng, H.W.; Chiu, J.H. Baicalein protects chicken embryonic cardiomyocyte against hypoxia–reoxygenation injury via μ-and δ-but not κ-opioid receptor signaling. Eur. J. Pharmacol. 2008, 588, 251–258. [Google Scholar] [CrossRef]
  181. Protić, D.; Beleslin-Čokić, B.; Spremović-Rađenović, S.; Radunović, N.; Heinle, H.; Šćepanović, R.; Gojković Bukarica, L. The Different Effects of Resveratrol and Naringenin on Isolated Human Umbilical Vein: The Role of ATP-Sensitive K+ Channels. Phytother. Res. 2014, 28, 1412–1418. [Google Scholar] [CrossRef]
  182. Yow, T.T.; Pera, E.; Absalom, N.; Heblinski, M.; Johnston, G.A.; Hanrahan, J.R.; Chebib, M. Naringin directly activates inwardly rectifying potassium channels at an overlapping binding site to tertiapin-Q. Br. J. Pharmacol. 2011, 163, 1017–1033. [Google Scholar] [CrossRef]
  183. Loscalzo, L.M.; Yow, T.T.; Wasowski, C.; Chebib, M.; Marder, M. Hesperidin induces antinociceptive effect in mice and its aglicone, hesperetin, binds to μ-opioid receptor and inhibits GIRK1/2 currents. Pharmacol. Biochem. Behav. 2011, 99, 333–341. [Google Scholar] [CrossRef]
  184. Hammadi, R.; Kúsz, N.; Mwangi, P.W.; Kulmány, Á.; Zupkó, I.; Orvos, P.; Tálosi, L.; Hohmann, J.; Vasas, A. Isolation and pharmacological investigation of compounds from Euphorbia matabelensis. Nat. Prod. Commun. 2019, 14, 1934578X19863509. [Google Scholar] [CrossRef] [Green Version]
  185. Lesage, F.; Lazdunski, M. Molecular and functional properties of two-pore-domain potassium channels. Am. J. Physiol. Ren. Physiol. 2000, 279, F793–F801. [Google Scholar] [CrossRef] [PubMed]
  186. Feliciangeli, S.; Chatelain, F.C.; Bichet, D.; Lesage, F. The family of K2P channels: Salient structural and functional properties. J. Physiol. 2015, 593, 2587–2603. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  187. Herrera-Pérez, S.; Campos-Ríos, A.; Rueda-Ruzafa, L.; Lamas, J.A. Contribution of K2P potassium channels to cardiac physiology and pathophysiology. Int. J. Mol. Sci. 2021, 22, 6635. [Google Scholar] [CrossRef]
  188. Wiedmann, F.; Frey, N.; Schmidt, C. Two-pore-domain potassium (K2P-) channels: Cardiac expression patterns and disease-specific remodelling processes. Cells 2021, 10, 2914. [Google Scholar] [CrossRef]
  189. Andres-Bilbe, A.; Castellanos, A.; Pujol-Coma, A.; Callejo, G.; Comes, N.; Gasull, X. The background K+ channel TRESK in sensory physiology and pain. Int. J. Mol. Sci. 2020, 21, 5206. [Google Scholar] [CrossRef] [PubMed]
  190. Luo, Y.; Huang, L.; Liao, P.; Jiang, R. Contribution of Neuronal and Glial Two-Pore-Domain Potassium Channels in Health and Neurological Disorders. Neural Plast. 2021, 2021, 8643129. [Google Scholar] [CrossRef]
  191. Djillani, A.; Mazella, J.; Heurteaux, C.; Borsotto, M. Role of TREK-1 in health and disease, focus on the central nervous system. Front. Pharmacol. 2019, 10, 379. [Google Scholar] [CrossRef] [Green Version]
  192. Medhurst, A.D.; Rennie, G.; Chapman, C.G.; Meadows, H.; Duckworth, M.D.; Kelsell, R.E.; Gloger, I.I.; Pangalos, M.N. Distribution analysis of human two pore domain potassium channels in tissues of the central nervous system and periphery. Mol. Brain Res. 2001, 86, 101–114. [Google Scholar] [CrossRef]
  193. Ren, K.; Liu, H.; Guo, B.; Li, R.; Mao, H.; Xue, Q.; Yao, H.; Wu, S.; Bai, Z.; Wang, W. Quercetin relieves D-amphetamine-induced manic-like behaviour through activating TREK-1 potassium channels in mice. Br. J. Pharmacol. 2021, 178, 3682–3695. [Google Scholar] [CrossRef]
  194. Kim, E.J.; Kang, D.; Han, J. Baicalein and wogonin are activators of rat TREK-2 two-pore domain K+ channel. Acta Physiol. 2011, 202, 185–192. [Google Scholar] [CrossRef] [PubMed]
  195. Gierten, J.; Ficker, E.; Bloehs, R.; Schlömer, K.; Kathöfer, S.; Scholz, E.; Zitron, E.; Katus, H.; Karle, C.; Thomas, D. Inhibition of hK2P3. 1 (TASK-1) Potassium Channels by the Tyrosine Kinase Inhibitor Genistein. Biophys. J. 2009, 96, 563a. [Google Scholar] [CrossRef] [Green Version]
  196. Gierten, J.; Ficker, E.; Bloehs, R.; Schlömer, K.; Kathöfer, S.; Scholz, E.; Zitron, E.; Kiesecker, C.; Bauer, A.; Becker, R.; et al. Regulation of two-pore-domain (K2P) potassium leak channels by the tyrosine kinase inhibitor genistein. Br. J. Pharmacol. 2008, 154, 1680–1690. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  197. Kirkegaard, S.S.; Lambert, I.H.; Gammeltoft, S.; Hoffmann, E.K. Activation of the TASK-2 channel after cell swelling is dependent on tyrosine phosphorylation. Am. J. Physiol. Cell Physiol. 2010, 299, C844–C853. [Google Scholar] [CrossRef] [PubMed]
  198. Fusi, F.; Trezza, A.; Tramaglino, M.; Sgaragli, G.; Saponara, S.; Spiga, O. The beneficial health effects of flavonoids on the cardiovascular system: Focus on K+ channels. Pharmacol. Res. 2020, 152, 104625. [Google Scholar] [CrossRef]
  199. He, J.; Li, S.; Ding, Y.; Tong, Y.; Li, X. Research Progress on Natural Products’ Therapeutic Effects on Atrial Fibrillation by Regulating Ion Channels. Cardiovasc. Ther. 2022, 2022, 4559809. [Google Scholar] [CrossRef]
  200. Scholz, E.P.; Zitron, E.; Katus, H.A.; Karle, C.A. Cardiovascular ion channels as a molecular target of flavonoids. Cardiovasc. Ther. 2010, 28, e46–e52. [Google Scholar] [CrossRef]
  201. Van Dijk, C.; Driessen, A.J.; Recourt, K. The uncoupling efficiency and affinity of flavonoids for vesicles. Biochem. Pharmacol. 2000, 60, 1593–1600. [Google Scholar] [CrossRef] [Green Version]
  202. Ulrih, N.P.; Ota, A.; Šentjurc, M.; Kure, S.; Abram, V. Flavonoids and cell membrane fluidity. Food Chem. 2010, 121, 78–84. [Google Scholar] [CrossRef]
  203. Selvaraj, S.; Krishnaswamy, S.; Devashya, V.; Sethuraman, S.; Krishnan, U.M. Influence of membrane lipid composition on flavonoid–membrane interactions: Implications on their biological activity. Prog. Lipid Res. 2015, 58, 1–13. [Google Scholar] [CrossRef]
  204. Erlejman, A.; Verstraeten, S.; Fraga, C.; Oteiza, P. The interaction of flavonoids with membranes: Potential determinant of flavonoid antioxidant effects. Free Radic. Res. 2004, 38, 1311–1320. [Google Scholar] [CrossRef] [PubMed]
  205. Hendrich, A.B. Flavonoid-membrane interactions: Possible consequences for biological effects of some polyphenolic compounds. Acta Pharmacol. Sin. 2006, 27, 27–40. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  206. Verstraeten, S.V.; Fraga, C.G.; Oteiza, P.I. Flavonoid–Membrane interactions: Consequences for biological actions. In Plant Phenolics and Human Health: Biochemistry, Nutrition, and Pharmacology; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2009; pp. 107–135. [Google Scholar]
  207. Veiko, A.G.; Sekowski, S.; Lapshina, E.A.; Wilczewska, A.Z.; Markiewicz, K.H.; Zamaraeva, M.; Zhao, H.c.; Zavodnik, I.B. Flavonoids modulate liposomal membrane structure, regulate mitochondrial membrane permeability and prevent erythrocyte oxidative damage. Biochim. Biophys. Acta (BBA) Biomembr. 2020, 1862, 183442. [Google Scholar] [CrossRef] [PubMed]
  208. Lee, A.G. How lipids affect the activities of integral membrane proteins. Biochim. Biophys. Acta (BBA) Biomembr. 2004, 1666, 62–87. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  209. Hedger, G.; Sansom, M.S. Lipid interaction sites on channels, transporters and receptors: Recent insights from molecular dynamics simulations. Biochim. Biophys. Acta (BBA) Biomembr. 2016, 1858, 2390–2400. [Google Scholar] [CrossRef]
  210. Corradi, V.; Sejdiu, B.I.; Mesa-Galloso, H.; Abdizadeh, H.; Noskov, S.Y.; Marrink, S.J.; Tieleman, D.P. Emerging diversity in lipid–protein interactions. Chem. Rev. 2019, 119, 5775–5848. [Google Scholar] [CrossRef] [Green Version]
  211. Levental, I.; Lyman, E. Regulation of membrane protein structure and function by their lipid nano-environment. Nat. Rev. Mol. Cell Biol. 2022, 1–16. [Google Scholar] [CrossRef]
  212. Yam, M.F.; Tan, C.S.; Shibao, R. Vasorelaxant effect of sinensetin via the NO/sGC/cGMP pathway and potassium and calcium channels. Hypertens. Res. 2018, 41, 787–797. [Google Scholar] [CrossRef]
  213. Jiang, H.; Xia, Q.; Wang, X.; Song, J.; Bruce, I. Luteolin induces vasorelaxion in rat thoracic aorta via calcium and potassium channels. Die-Pharm. Int. J. Pharm. Sci. 2005, 60, 444–447. [Google Scholar]
  214. Lee, W.; Woo, E.R.; Lee, D.G. Effect of apigenin isolated from Aster yomena against Candida albicans: Apigenin-triggered apoptotic pathway regulated by mitochondrial calcium signaling. J. Ethnopharmacol. 2019, 231, 19–28. [Google Scholar] [CrossRef]
  215. Lee, J.H.; Amarsanaa, K.; Wu, J.; Jeon, S.C.; Cui, Y.; Jung, S.C.; Park, D.B.; Kim, S.J.; Han, S.H.; Kim, H.W.; et al. Nobiletin attenuates neurotoxic mitochondrial calcium overload through K+ influx and ΔΨm across mitochondrial inner membrane. Korean J. Physiol. Pharmacol. Off. J. Korean Physiol. Soc. Korean Soc. Pharmacol. 2018, 22, 311. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  216. Donato, F.; Filho, C.B.; Giacomeli, R.; Alvater, E.E.T.; Fabbro, L.D.; Antunes, M.d.S.; de Gomes, M.G.; Goes, A.T.R.; Souza, L.C.; Boeira, S.P.; et al. Evidence for the involvement of potassium channel inhibition in the antidepressant-like effects of hesperidin in the tail suspension test in mice. J. Med. Food 2015, 18, 818–823. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  217. Roohbakhsh, A.; Parhiz, H.; Soltani, F.; Rezaee, R.; Iranshahi, M. Molecular mechanisms behind the biological effects of hesperidin and hesperetin for the prevention of cancer and cardiovascular diseases. Life Sci. 2015, 124, 64–74. [Google Scholar] [CrossRef] [PubMed]
  218. Ngo, Q.A.; Thi, T.H.N.; Pham, M.Q.; Delfino, D.; Do, T.T. Antiproliferative and antiinflammatory coxib–Combretastatin hybrids suppress cell cycle progression and induce apoptosis of MCF7 breast cancer cells. Mol. Divers. 2021, 25, 2307–2319. [Google Scholar] [CrossRef]
  219. Al-Ishaq, R.K.; Abotaleb, M.; Kubatka, P.; Kajo, K.; Büsselberg, D. Flavonoids and their anti-diabetic effects: Cellular mechanisms and effects to improve blood sugar levels. Biomolecules 2019, 9, 430. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Den Hartogh, D.J.; Tsiani, E. Antidiabetic properties of naringenin: A citrus fruit polyphenol. Biomolecules 2019, 9, 99. [Google Scholar] [CrossRef] [Green Version]
  221. Li, S.; Zhang, Y.; Sun, Y.; Zhang, G.; Bai, J.; Guo, J.; Su, X.; Du, H.; Cao, X.; Yang, J.; et al. Naringenin improves insulin sensitivity in gestational diabetes mellitus mice through AMPK. Nutr. Diabetes 2019, 9, 28. [Google Scholar] [CrossRef] [Green Version]
  222. Multhoff, G.; Molls, M.; Radons, J. Chronic inflammation in cancer development. Front. Immunol. 2012, 2, 98. [Google Scholar] [CrossRef] [Green Version]
  223. Michels, N.; van Aart, C.; Morisse, J.; Mullee, A.; Huybrechts, I. Chronic inflammation towards cancer incidence: A systematic review and meta-analysis of epidemiological studies. Crit. Rev. Oncol. 2021, 157, 103177. [Google Scholar] [CrossRef]
  224. Iervolino, M.; Lepore, E.; Forte, G.; Laganà, A.S.; Buzzaccarini, G.; Unfer, V. Natural molecules in the management of polycystic ovary syndrome (PCOS): An analytical review. Nutrients 2021, 13, 1677. [Google Scholar] [CrossRef]
  225. Wawrzkiewicz-Jałowiecka, A.; Kowalczyk, K.; Trybek, P.; Jarosz, T.; Radosz, P.; Setlak, M.; Madej, P. In Search of New Therapeutics—Molecular Aspects of the PCOS Pathophysiology: Genetics, Hormones, Metabolism and Beyond. Int. J. Mol. Sci. 2020, 21, 7054. [Google Scholar] [CrossRef] [PubMed]
  226. Wawrzkiewicz-Jałowiecka, A.; Lalik, A.; Soveral, G. Recent Update on the molecular mechanisms of gonadal steroids action in adipose tissue. Int. J. Mol. Sci. 2021, 22, 5226. [Google Scholar] [CrossRef] [PubMed]
  227. Ceccarelli, I.; Bioletti, L.; Peparini, S.; Solomita, E.; Ricci, C.; Casini, I.; Miceli, E.; Aloisi, A.M. Estrogens and phytoestrogens in body functions. Neurosci. Biobehav. Rev. 2021, 132, 648–663. [Google Scholar] [CrossRef]
  228. Gorzkiewicz, J.; Bartosz, G.; Sadowska-Bartosz, I. The potential effects of phytoestrogens: The role in neuroprotection. Molecules 2021, 26, 2954. [Google Scholar] [CrossRef] [PubMed]
  229. Torrens-Mas, M.; Roca, P. Phytoestrogens for cancer prevention and treatment. Biology 2020, 9, 427. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The schematic summary of the impact of key representatives of the main groups of flavonoids on the activity of the BK channels. Arrow up corresponds to the increase of the open state probability. Arrow down denotes channel inhibition. Both arrows represent the case when different types of channel modulation were reported depending on the cell types where the investigated BK channels were expressed.
Figure 1. The schematic summary of the impact of key representatives of the main groups of flavonoids on the activity of the BK channels. Arrow up corresponds to the increase of the open state probability. Arrow down denotes channel inhibition. Both arrows represent the case when different types of channel modulation were reported depending on the cell types where the investigated BK channels were expressed.
Ijms 24 01311 g001
Figure 2. A graphical representation of the selected important biological processes that can be affected by flavonoid administration, for which the molecular control mechanism can be, at least, partially explained by the K + channel modulation (with appropriate examples). The ↑ represents channels’ activation and ↓ denotes channels’ inhibition.
Figure 2. A graphical representation of the selected important biological processes that can be affected by flavonoid administration, for which the molecular control mechanism can be, at least, partially explained by the K + channel modulation (with appropriate examples). The ↑ represents channels’ activation and ↓ denotes channels’ inhibition.
Ijms 24 01311 g002
Table 1. Effects of flavonoids on the activity of different subtypes of the Kv channels. HTL stands for human T lymphocytes. HLTJ is human leukemic T cells. E C 50 is defined as the concentration of a flavonoid that gives the half-maximal response. I C 50 is the concentration of a flavonoid concentration at 50% channel inhibition. The arrows symbolize the type of observed effects on the channel activity: ↓ inhibition, ↑ activation, → no effect. As an exception, the table includes resveratrol, which is not a flavonoid, but it was analyzed in the same series of experiments as flavonoids.
Table 1. Effects of flavonoids on the activity of different subtypes of the Kv channels. HTL stands for human T lymphocytes. HLTJ is human leukemic T cells. E C 50 is defined as the concentration of a flavonoid that gives the half-maximal response. I C 50 is the concentration of a flavonoid concentration at 50% channel inhibition. The arrows symbolize the type of observed effects on the channel activity: ↓ inhibition, ↑ activation, → no effect. As an exception, the table includes resveratrol, which is not a flavonoid, but it was analyzed in the same series of experiments as flavonoids.
Kv Channel SubtypeFlavonoidType of CellEffect IC 50 / EC 50 References
Kv1.3GenisteinHTL30–60 μ M Teisseyre et al. [25] (2005)
DaidzeinHTL Teisseyre et al. [25] (2005)
6-PrenylnaringeninHLJT 5.8 μ M Teisseyre et al. [26] (2018)
AcacetinHLJT 30 μ M Teisseyre et al. [26] (2018)
21 μ M Zhao et al. [27] (2014)
ChrysinHLJT26 μ M Teisseyre et al. [26] (2018)
Chrysin + mevastatinHLTJ8 μ M Teisseyere et al. [28] (2022)
Chrysin + simvastatinHLTJ 11 μ M Teisseyere et al. [28] (2022)
BaicaleinHLJT Teisseyre et al. [26] (2018)
WogoninHLJT Teisseyre et al. [26] (2018)
LuteolinHLJT Teisseyre et al. [26] (2018)
ResveratrolHTL 41 μ M Teisseyre et al. [29] (2006)
NaringeninHTL Teisseyre et al. [30] (2009)
Naringenin-4 ,7-dimethyletherHTL Teisseyre et al. [30] (2009)
HLTJ 13 μ M Gąsiorowska et al. [31] (2015)
Naringenin-7-methyletherHTL Teisseyre et al. [30] (2009)
HLJT 16 μ M Gąsiorowska et al. [31] (2009)
AromadendrinHTL Teisseyre et al. [30] (2009)
IsoxanthohumolHLJT 7.8 μ M Gąsiorowska et al. [31] (2015)
XanthohumolHLJT 3.1 μ M Gąsiorowska et al. [31] (2015)
8-prenylnaringeninHLJT Gasiorowska et al. [32] (2012)
Licochalcone AHLJT 0.83 μ M Phan et al. [33] (2021)
8-prenylnaringenin+mevastatinHLTJ 7 μ M Teisseyre et al. [28] (2022)
Kv1.5MyricetinHEK 293 Ou et al. [34] (2016)
HesperetinHEK 293 23 μ M Wang et al. [35] (2016)
QuercetinXenopus oocytes 37.8 μ M Yang et al. [36] (2009)
rats (in vivo) Morales-Cano et al. [37] (2014)
HEK 293 Liu et al. [38] (2012)
3,7,3 ,4 -tetramethylquecertinHEK 293 Liu et al. [38] (2012)
3,5,7,3 ,4 -pentamethylquecertinHEK 293 Liu et al. [38] (2012)
ApigeninHEK 293 Liu et al. [38] (2012)
7,4 -dimethylapigeninHEK 293 Liu et al. [38] (2012)
5,7,4 -trimethylapigeninHEK 293 6.4 μ M Liu et al. [38] (2012)
EGCGCHO 101 μ M Choi et al. [39] (2001)
IsoliquiritigeninH9c2 Noguchi et al. [40] (2008)
AcacetinHEK 293 Wu et.al [41] (2011)
atrial myocytes 3.2 μ M Li et al. [42] (2008)
Kv1.7PuerarinHEK 293 36 μ M Xu et al. [43] (2016)
(−)-Epigallocatechin-3-gallateCHO 30 μ M Kang et al. [44]
NaringeninCHO 110 μ M Sanson et al. [45] (2022)
Kv2.1IsoliquiritigeninH9c2 0.11 μ M Noguchi et al. [40] (2008)
GenisteinHEK 293 Aréchiga-Figueroa et al. [46] (2017)
Naringenin-4 ,7-dimethyletherCHO 21 μ M Gu et.al [47] (2022)
Kv4.3GenisteinCHO 125 μ M Kim et al. [48] (2011)
DaidzeinCHO Kim et al. [48] (2011)
GenistinCHO Kim et al. [48] (2011)
Epigallocatechin-3-gallateCHO Kang et al. [44] (2010)
NaringeninCHO 115 μ M Sanson et al. [45] (2022)
5,7,4 -trimethylapigeninhuman atrial myocytes 19.8 μ M Liu et al. [38] (2012)
Kv10.1Procyanidin B1HEK 10 μ M Na et al. [49] (2020)
Table 3. The effects of different flavonoids on the activity of BK channels’ isoforms in different cell types. The ↑ represents channel activation, and ↓ denotes channel inhibition. The table includes the effects of berberine, which does not strictly belong to the flavonoid family. Nevertheless, by some authors, it is categorized as ’isoquinoline flavonoid’.
Table 3. The effects of different flavonoids on the activity of BK channels’ isoforms in different cell types. The ↑ represents channel activation, and ↓ denotes channel inhibition. The table includes the effects of berberine, which does not strictly belong to the flavonoid family. Nevertheless, by some authors, it is categorized as ’isoquinoline flavonoid’.
FlavonoidMaterialEffectReferences
Naringeninrat aortic ringsSaponara et al. [102] (2006)
HEK 293THsu et al. [103] (2014)
colonic smooth muscle cellsYang et al. [104] (2014)
rat tracheal smooth muscle cellsShi et al. [105] (2019)
mitoplasts from rat heart (left ventricular tissue)Tesai et al. [106,107] (2013, 2017)
mitoplasts from primary human dermal fibroblastsKampa et al. [108] (2019)
mitoplasts from human endothelial cells EA.hy926Kicinska et al. [109] (2020)
Naringinrat tracheal smooth muscle cellsShi et al. [105] (2019)
Diocleinrat small mesenteric arteriesCortes et al. [112] (2001)
Hesperidinrat hippocampal cellsDimpfel et al. [113] (2006)
Hesperetinrat hippocampal cellsDimpfel et al. [113] (2006)
Quercetinhuman bladder cancer cellsKim et al. [114] (2011)
murine smooth muscles (ileal myocytes)Melnyk et al. [115] (2019)
rat coronary smooth muscle cellsZhang et al. [116] (2020)
mitoplasts from human endothelial cells EA.hy926Kampa et al. [117,118] (2021, 2022)
KaempferolXenopus oocytesLi et al. [119] (1997)
human umbilical vein endothelial cellsXu et al. [121] (2008)
porcine coronary arteryXu et al. [122] (2015)
rat pulmonary arteryMahobiya et al. [120] (2018)
Luteolinmitoplasts from rat cardiomyocytes, mitoplasts from human endothelial cells EA.hy926Kampa et al. [123] (2022)
Baicaleinrat tracheal smooth muscleSaadat et al. [125] (2019)
ApigeninXenopus oocytesLi et al. [119] (1997)
Morelloflavonerat thoracic aortaLamai et al. [126] (2013)
Genisteinrat vascular smooth muscle cellsBai et al. [127] (2020)
vascular smooth muscle cellsSun et al. [128] (2015)
rat superior mesenteric arteryZhou et al. [129] (2005)
HEK 293 cellsWang et al. [130] (2017)
rat mesenteric artery ringsNevala et al. [133] (2001)
bovine trabecular meshwork cellsStumpff et al. [131] (1999)
Daidzeinrat cerebral basilar artery smooth muscle cellsZhang et al. [132] (2010)
Xenopus oocytesSun et al. [90] (2007)
rat mesenteric artery ringsNevala et al. [133] (2001)
PuerarinXenopus oocytesSun et al. [90] (2007)
mitochondria of rat cardiomyocytesYang et al. [134] (2008)
mitochondria of rat cardiomyocytesYao et al. [135] (2010)
Phloretinmyelinated nerve fibres of Xenopus laevisKoh et al. [136] (1994)
heterologous expression models (unspecified in the cited work)Gonzalez et al. [137,138] (2012, 2013)
Nothofaginrat kidney cellsMarques et al. [139] (2020)
Berberinecerebral vascular smooth muscle cellsMa et al. [141] (2017)
Rottlerinmurine tracheal smooth muscleGoldklang et al. [144] (2013)
human hepatic stellate cellsYang et al. [145] (2020)
Table 4. The effects of flavonoids on the KATP channels. The ↑ represents channel activation and ↓ denotes channel inhibition, while → stands for no effect on channel activity.
Table 4. The effects of flavonoids on the KATP channels. The ↑ represents channel activation and ↓ denotes channel inhibition, while → stands for no effect on channel activity.
KATP ChannelsFlavonoidCell TypeEffectReferences
Kir6.1QuercetinRat norvegicus aorta/MDTrezza et al. [159] (2018)
5–HydroxyflavoneRat norvegicus aorta/MDTrezza et al. [159] (2018)
isovitexinrat isolated mesenteric bedsTirloni et al. [154] (2019)
Kir6.2CyanidinRat Pancreatic β -cells INS-1Suantawee et al. [169] (2017)
(−)-Epigallocatechin-3-gallateXenopus oocytesJin et al. [162] (2007)
(−)-Epicatechin-3-gallateXenopus oocytesJin et al. [162] (2007)
(−)-EpicatechineXenopus oocytesJin et al. [162] (2007)
(−)-EpigallocatechinXenopus oocytesJin et al. [162] (2007)
Kir6.xBerberineGuinea pig ventricular myocytesHua Z et al. [167] (1994)
Kir6.xBerberineGuinea pig ventricular myocytesWang et al. [168] (1996)
Kir6.xNaringeninmyocardial cells of Sprague-Dawley ratsMeng et al. [166] (2016)
Kir6.xNaringeninHuman Umbilical VeinProtic et al. [181] (2014)
Kir6.xBaicelinRat tracheal smooth muscleSaadat et al. [125] (2019)
Kir6.xTFRGat cardiomyocytesJiao Li et.al. [157] (2015)
Kir6.xGenisteinRabbit portal vein smooth muscleOgata et.al. [174] (1997)
Kir6.xBaicaleinMice gastric mucosal ulcerationsRibeiro et.al. [170] (2016)
Kir6.xMorinMitoplasts from rat myocardial cellsRameshrad et.al. [160] (2021)
Kir6.xChrysoeriolrabbit jejunum and aortic rings, guinea-pig tracheaKhan et.al. [161] (2006)
Kir6.xVitexinrabbit jejunum, guinea-pig tracheaKhan et.al. [161] (2006)
Kir6.xOrientinrabbit jejunumKhan et.al. [161] (2006)
mitoKATPPuerarinRat cardiomyocytesYao et al. [135] (2012)
mitoKATPNaringeninRat cardiomyocytesMeng et al. [166] (2016)
mitoKATPBaicaleinChicken embryonic cardiomyocyteTu et al. [180] (2008)
mitoKATP(−)-Epigallocatechin-3-gallateRat cardiomyocytesSong et al. [176] (2010)
mitoKATPTheaflavinRat cardiomyocytesMa et al. [177] (2011)
mitoKATPProanthocyanidinsRat cardiomyocytesHu et al. [178] (2014)
mitoKATPGenisteinRabbit cardiomyocytesYao et al. [179] (2009)
mitoKATPMorinRat cardiomyocytesYao et al. [160] (2021)
Table 5. The effects of flavonoids on the GIRK channels. The ↑ represents channel activation and ↓ denotes channel inhibition.
Table 5. The effects of flavonoids on the GIRK channels. The ↑ represents channel activation and ↓ denotes channel inhibition.
GIRK ChannelsFlavonoidCell TypeEffectReferences
Kir3.1/Kir3.2HesperidinXenopus laevis oocytesLoscalzo et al. [183] (2011)
Kir3.1/Kir3.4EriodictyolHEK-293 (human embryonic kidney)Hammadi et al. [184] (2019)
Kir3NaringinXenopus laevis oocytesYow et al. [182] (2011)
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

Richter-Laskowska, M.; Trybek, P.; Delfino, D.V.; Wawrzkiewicz-Jałowiecka, A. Flavonoids as Modulators of Potassium Channels. Int. J. Mol. Sci. 2023, 24, 1311. https://doi.org/10.3390/ijms24021311

AMA Style

Richter-Laskowska M, Trybek P, Delfino DV, Wawrzkiewicz-Jałowiecka A. Flavonoids as Modulators of Potassium Channels. International Journal of Molecular Sciences. 2023; 24(2):1311. https://doi.org/10.3390/ijms24021311

Chicago/Turabian Style

Richter-Laskowska, Monika, Paulina Trybek, Domenico Vittorio Delfino, and Agata Wawrzkiewicz-Jałowiecka. 2023. "Flavonoids as Modulators of Potassium Channels" International Journal of Molecular Sciences 24, no. 2: 1311. https://doi.org/10.3390/ijms24021311

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