Abstract
Background:
Ion channels play crucial roles in cellular biology, physiology, and communication including sensory perception. Voltage-gated potassium (Kv) channels execute their function by sensor activation, pore-coupling, and pore opening leading to K+ conductance.
Scope of review:
This review focuses on a voltage-gated K+ ion channel KCNQ1 (Kv 7.1). Firstly, discussing its positioning in the human ion chanome, and the role of KCNQ1 in the multitude of cellular processes. Next, we discuss the overall channel architecture and current structural insights on KCNQ1. Finally, the gating mechanism involving members of the KCNE family and its interaction with non-KCNE partners.
Major conclusions:
KCNQ1 executes its important physiological functions via interacting with KCNE1 and non-KCNE1 proteins/molecules: calmodulin, PIP2, PKA. Although, KCNQ1 has been studied in great detail, several aspects of the channel structure and function still remain unexplored. This review emphasizes the structural and biophysical studies of KCNQ1, its interaction with KCNE1 and non-KCNE1 proteins and focuses on several seminal findings showing the role of VSD and the pore domain in the channel activation and gating properties.
General significance:
KCNQ1 mutations can result in channel defects and lead to several diseases including atrial fibrillation and long QT syndrome. Therefore, a thorough structure-function understanding of this channel complex is essential to understand its role in both normal and disease biology. Moreover, unraveling the molecular mechanisms underlying the regulation of this channel complex will help to find therapeutic strategies for several diseases.
Keywords: Chanome, Channelopathy, Potassium ion channel, KCNQ1, KCNE1, Long QT syndrome
Graphical abstract

1. Introduction
Ion channels are a ubiquitous class of pore-forming proteins that facilitate the passage of ions across the lipid bilayer in multicellular organisms. Myriad crucial physiological and biological functions are modulated by ion channels, including neurotransmitter release, muscle contraction, electrolyte balance, hormone secretion, sensory transduction, and cognition [1–5].
The classification of ion channels is based on the nature of the gating stimulus, which could be transmembrane voltage potential, chemical ligands, light, temperature, secondary messengers, or mechanical. Yu and Catterall (2004), proposed the term “Chanome” comprising all ion channel proteins, similar to “kinome” for the protein kinase superfamily1 (Figure 1 A). The chanome was further sub-grouped into a VGL-chanome: the superfamily of voltage-gated-like ion channels, and a LG-chanome: ligand-gated ion channels. Ion channels that execute passive ion flux driven by an electrochemical gradient and are regulated by binding of small molecule ligands are called ligand-gated ion channels (LGICs) [1–3]. The gating process of LGIC conductance involves the binding of specific ligands and subsequent conformational changes leading to the opening of the channel. LGICs play important roles in the somatic neuromuscular junction and fast synaptic transmission and include glycine receptors, P2X receptors, the nicotinic acetylcholine receptor (nAchR), ionotropic glutamate (NMDA, AMPA, and Kainate receptors), 5-HT3, GABA, IP3 receptor, acid-sensing (proton-gated) ion channels (ASICs), epithelial sodium channels (ENaC), and zinc-activated channels (ZAC) [1,4–7]. Voltage-gated ion channels are regulated by changes in the transmembrane voltage potential and classified as „voltage-gated-like (VGL) ion channel chanome‟, which includes the voltage-gated potassium ion channel superfamily [1,3–6] (Kv superfamily, Figure 1B). VGLs consists of distinct structural domains with transmembrane helices S1–S4, comprising the voltage sensor domain (VSD), and S5–S6, forming the pore domain (PD) [7]. The general gating mechanism of VGLs involves movement of the S4 transmembrane helix of VSD towards the extracellular region. The voltage sensor in S4 helix is lined by several basic amino acid residues that sense the change in membrane potential. The S4 helix then translates toward the extracellular milieu, resulting in structural changes that are propagated to the PD via the S4–S5 linker and/ by the P-loop region between S5 and S6, which contains the selectivity filter [7,8].
Figure 1.
(A) Ion-channel chanome depicting voltage-gated ion channel superfamily with different sub-classes. (B) Floral diagram depiction of human potassium ion channel (Kv) superfamily and the relative position of KCNQ1 (Kv 7.1) in the channel superfamily.
Heritable mutations in the genes that encode ion channels can result in channel dysfunction or loss of function, causing diseases collectively termed as channelopathies [4–6, 9–13]. Aberrant loss of function gain of function, or dysfunction can result in diseases like epilepsy, asthma, cystic fibrosis, myasthenia gravis, cancer, cardiac arrhythmia, diabetes, deafness, migraine, blindness, hypertension, and multiple sclerosis [4–6, 9–13]. Ion channels have been successfully targeted with several classes of drugs. For example, volatile anesthetic drugs primarily block LGICs. Gamma-aminobutyric acid analogs (gabapentinoids) are used to treat epilepsy and anxiety disorder. Calcium channel blockers (CCBs) are used to treat hypertension patients, whereas, sodium channel blockers are primarily used to treat epilepsy [4–6, 9–13]. IKS channel blockers are a promising class of antiarrhythmic drugs. Chromanol 293B is an IKS channel blocker responsible for the prolongation of cardiac action potential. Lerche et al. (2000) showed the inner pore domain of KCNQ1 to be a molecular target of Chromanol 293B. They also identified amino acid mutations V307L and I337V, located in the evolutionary conserved pore loop (PD) region and the S6 transmembrane domain of KCNQ1 almost completely inhibited Chromanol 293B effect [14]. Same effect was observed with mutant KCNQ1/E1 channel. However, mutations involving KCNE1 in the KCNQ1/E1 complex, and its effect on inhibition by Chromanol 293B revealed no direct involvement of KCNE1 subunit as demonstrated by IKS current measurements. It is asserted that KCNE1 acts allosterically to alter the conformation of KCNQ1to modulate the channel properties [14]. Similar findings were reported by Xu et al., where KCNE1 induced backbone displacement in the selectivity filter region (T312) of KCNQ1. However, no direct contact between KCNE1 and KCNQ1 PD was observed indicating allosteric modulation of the KCNQ1 selectivity filter in the presence of KCNE1 [13].
Several drugs elicit cardiotoxic effects by binding to the KCNQ1/E1 complex, blocking the conduction of K+ ions and disturbing the generation of IKS currents [15]. Fluoxetine, an antidepressant drug has been shown to induce LQTS at high serum levels [15]. It acts by blocking IKr and also inhibiting the protein trafficking of IKr channel, thereby further reducing IKr. Fluoxetine and its derivative norfluoxetine also inhibit KCNQ1/E1 currents leading to QT prolongation. In the presence of K422T mutation in KCNQ1, QT prolongation lasted for 154 ms. Celecoxib, a non-specific Kv blocker is known to block KCNQ1 when coexpressed with KCNE1 subunits [16]. However, previous studies show that celecoxib acts as a KCNQ5 channel opener [17]. The mechanism underlying this opposite regulation remains elusive.
Channel proteins and their functional complexes have been studied using various biophysical techniques [18–27]. These include structural studies that have been instrumental in illuminating the functional properties of different ion channels and their channelopathic landscapes. This review is focused on KCNQ1, a voltage-gated potassium ion channel that has been of much interest over the past three decades because of its direct role in the cardiovascular diseases such as atrial fibrillation (AF) and long QT syndrome (LQTS) [10–13]. In this review, we provide a comprehensive and up-to-date understanding of the structure and function of KCNQ1 and also comment on future direction of this research field.
Functional versatility of KCNQ1 and ‘KCNQ1 channelopathy’
Biological membranes, as well as membrane proteins play crucial roles in normal and disease biology and therefore have been studied in the great detail [28–30]. Ion channels are a distinct class of membrane proteins regulating key physiological processes in the human body (Figure 2 A, B). KCNQ1 is expressed ubiquitously [31–33] and is the pore-forming subunit of the slow delayed rectifier current (IKS) channel complex [31–47] (Figure 2 A). The functional role of KCNQ1 is thoroughly studied both in the heart [34–37] and in the inner ear [10,13,38]. KCNQ1 forms a complex with the β-subunit KCNE1 (MinK) in the cardiac IKS channel to regulate the cardiac action potential duration [34–37]. The outward K+ current generated by KCNQ1/KCNE1 complex in the IKS channel in heart, is one of the repolarizing K+ currents that contribute to the cessation of the cardiac action potential. Defective cardiac IKS channels (originating from dysfunctional KCNQ1 protein) cause pathological changes in the action potential duration of the heart, leading to severe cardiac arrhythmias and long QT syndrome (due to a prolonged cardiac action potential duration) [10–13].
Figure 2.
(A) Physiological roles of KCNQ1 in human body (Figure design inspired from Liin et al.; ref. 33). (B) mutated or dysfunctional KCNQ1 is associated with various diseases.
The role of the KCNQ1/KCNE1 complex has been thoroughly studied in ear biology [10,13,38]. Inner ear endolymph K+ homeostasis is maintained by the K+ flux through IKS channel (KCNQ1/KCNE1 complex) generating an endocochlear potential [10,13,38]. Dysfunction of IKS channel results in decreased hearing or congenital deafness [13]. Mutations in KCNQ1 or KCNE1 are associated with two forms of Long QT syndrome [39]. Romano Ward syndrome, an autosomal dominant disorder and Jervell and Lange-Nielson syndrome, inherited as an autosomal recessive trait. The latter being less common but with severe sensorineural hearing loss (SNHL) [39]. The pathophysiology of this hearing impairment along with Long QT syndrome in Jervell and Lange-Nielson syndrome remains unclear. Besides heart and ear, KCNQ1 has been reported to form a channel complex with either KCNE1 or other members of KCNE family (E1–E5) members in the pancreas, kidney, colon and intestine, stomach, thyroid gland and airways [40–48] (Figure 2 A). Similar to the ear, the major function of the KCNQ1/KCNE complexes in these organs, is to contribute towards K+ homeostasis. For example, the K CNQ1/KCNE3 complex also plays a significant role in the maintenance of the proper membrane potential for transport of ions across the epithelium, for example, Cl− and gastric acid [42,45,47,48] (Figure 2 A). In the intestine, the KCNQ1/KCNE3 complex regulates Cl− ion secretion in a cAMP dependent manner. This is achieved by constitutively activating the open state of the KCNQ1/KCNE3 channel complex, that aids in the transepithelial transport of Cl− ions via basolateral recycling of K+ ions and by increasing the driving force generated by K+ efflux for Cl− secretion [47]. KCNQ1 affects the endocrinological functions by regulating the physiology of thyroid gland. KCNQ1/KCNE2 form a constitutively active K+ channel in the thyrocytes required for normal thyroid hormone biosynthesis [49]. The brain expresses KCNQ1 where it contributes towards neuronal excitability [11,12]. Major reported disease groups from dysfunctional KCNQ1 are long and short QT syndrome, Romano Ward syndrome, Gestational diabetes, Familial atrial fibrillation, and Jervell and Lange-Nielsen syndrome (Figure 2 B). Bartos et al. identified an R231H mutation in KCNQ1 that is responsible for interfamilial early onset of Atrial Fibrillation (AF) [50]. R231H variants showed increased KCNQ1 current (IKCNQ1) resulting in shortened atrial action potential (AP) duration. R231H was also found to be resistant towards PKA stimulation, suggesting a potential role of PKA in regulating IKCNQ1 in these variants [50]. Based on the reported expression of KCNQ1–KCNE complexes in various body tissues with different functions, the multi-functional aspects of this channel protein in normal and disease physiology is obvious.
2. KCNQ1 channel architecture and structural insights
The functional unit of KCNQ1 consists of a tetramer with each monomer containing a short N-terminal S0 helix, followed by six transmembrane (S1–S6) helices and four intracellular helices in the C-terminal domain [51–53] (Figure 3). Similar to canonical VGLs, helices S5 and S6 ofKCNQ1 line the interior of the complex in the tetramer forming the pore domain (PD). The PD includes the signature GYGD sequence and assists in K+ conductance across the channel [51–56]. The topology of the VSD is similar to other membrane-spanning 4 -helical bundle proteins found in nature [57–59]. However, several key structural features of KCNQ1 present a sharp contrast with not only with other members of the Kv superfamily but with all of the other VGLs characterized so far, making it unique with respect to voltage dependence and gating properties. Sun et al. (2017) pin-pointed two key structural differences in the VSD region of KCNQ1 (Q1-VSD) from Xenopus laevis (which shares 78% sequence identity with human KCNQ1) and other VSDs of Kv channels; 1) a nine amino-acid α-helical insertion in the S2–S3 linker region, and 2) a straight S3 helix throughout its length due to the lack of a conserved proline residue present in most Kv channels but absent in KCNQ1 [53]. KCNQ1 presents a unique functional feature when compared to other VGLs, where it can switch from the voltage dependent state to the constitutively active state. The voltage sensor domain (Q1-VSD) can lose its voltage dependence upon association with members of the KCNE family making it constitutively active [35].
Figure 3.
Topological depiction of KCNQ1 in a membrane environment.
Several basic amino acid residues line the S4 helix of Q1-VSD [53,56,60]. Upon membrane depolarization, these positively charged residues sense the change in membrane potential causing the S4 helix to move towards the extracellular region [53,56,60] (Figure 4A). This is accomplished by electrostatic interactions between basic residues residing in the S4 region and acidic residues called “gating charge transfer residues” within the S2 and S3 helical regions [53,61].
Figure 4.
(A) Membrane topology of KCNQ1-Voltage Sensor Domain (Q1-VSD) showing the movement of S4 helix during VSD activation. (B) Important role played by KCNQ1–S6 and S4 helices and their interaction with KCNE1 and KCNE3 (depited as E1 and E3 (yellow) in 4B and 4C respectively) proteins in the opening/closing (gating mechanism) of the channel complex.
A recent cryo-EM structure of KCNQ1 by Sun et al. [53] revealed the structural details underlying the decoupled-depolarized Q1-VSD conformation. While the pore of the channel in this structure is closed, the voltage sensor domain is in the fully activated state. They reported crucial positioning of F157 residue with respect to the gating charges that cap the charge transfer center in the depolarized conformation of Q1-VSD. Another distinct feature, suggesting the uncoupled VSD-PD conformation in the case of KCNQ1, is a structural perturbation that involves shift of a loop by ~10.5 Å near residues 234–236 along with the S4–S5 linker facing away from the S6 helix compared to Kv1.2–2.1. The loop region connecting the S5–S6 in KCNQ1 is lined by negatively charged amino-acids while S6 lines the inner pore of the channel. KCNQ1 does not have a conserved PXP motif (a salient feature of Kv channels) but it is replaced by a functionally significant PAG motif, which is involved in channel activation [60,62]. The soluble C-terminal region consists of four helices (HA-HD). Helices HA and HB interact with Calmodulin (CaM), an important player in the channel function and disease biology [53]. Helices HC and HD are thought to assist in the tetramerization of the channel [53].
One of the essential components of a functional KCNQ1 complex in the heart is its association with the KCNE1 subunit (minK or β) (Figure 5 A). KCNE family members (KCNE1–5) are single-pass transmembrane proteins which associate with KCNQ1 to form the channel complex. Previous studies suggest that up to 4 KCNE subunits could associate with KCNQ1 tetramer [63]. However, the optimal IKS is generated when there is 4:2 ratio between KCNQ1-KCNE1 [64]. Cysteine cross-linking studies have shown that KCNE1 aligns itself in the cleft between the two Q1-VSDs in the channel complex [65,66] (Figure 5 B). Association with KCNE1 slows down the activation kinetics of IKS which is crucial during the repolarization phase of cardiac action potential [34,67].
Figure 5.
(A) KCNQ1-KCNE1 channel complex, where each subunit of KCNQ1 is labeled as 1–4 and S1–S6 helices of KCNQ1 are shown in different colors. (B) Top cartoon view of the complex showing VSD in the cleft between KCNQ1 and KCNE1 (shown as E1 in red) monomers. The pore domain is surrounded by blue colored S5–S6 helices.
Hasani et al. used molecular dynamic (MD) simulations and compared un-complexed KCNQ1 and KCNQ1/KCNE1 with respect to K+ ion transport across the membrane [68]. They reported four binding sites for K+ ions in the selectivity filter region (TIGYG) of KCNQ1 protein that direct its passage through the filter. The presence of KCNE1 causes constriction of the pore, slowing down the passage of K+ ions without causing significant changes to the selectivity filter region, thereby slowing down the activation kinetics of KCNQ1, as reported by Xu et al [69]. As opposed to previous studies, Xu et al. found that side chain interactions predominantly dictated the interactions between KCNQ1 and KCNE1 in the extracellular region, which impacted the voltage dependent activation of the channel [69]. This change could be attributed to the inward movement of S6 helical region in the presence of KCNE1.
3. KCNQ1 voltage dependent gating mechanism
Similar to other Kv channels, KCNQ1 follows a sequential, voltage dependent three step activation, switching from the initial resting state to an intermediate state followed by a fully activated channel [70,71]. We recently reported the topology of the isolated Q1-VSD in the lipid bilayer in its intermediate state [72]. KCNQ1 is unique in the fact that the pore remains open in both the intermediate as well as in the activated state of the Q1-VSD. However, both states demonstrate different gating and permeation properties indicating that the PD-VSD interactions play a role in determining the channel conformation and K+ transport.
The S4 helix of Q1-VSD plays an important role in the activation of the voltage sensor. This is assisted by several conserved arginine residues localized in S4 [73]. KCNQ1 differs from other canonical Kv channels in that it has fewer basic amino-acid residues lining the S4 helix. Previous studies have shown that Q1-VSD activation is a two-step process involving interactions between E160 in S2 and S4 arginines [74]. Charge reversal mutations revealed that while the E160R mutation caused loss of current in voltage-clamp fluorometry (VCF) experiments, charge reversal at R231(R231E) or R237 (R237E) partially restored it indicating the arrest of VSD in the states where these residues interact. It was also observed that the mutant pairs E160R/R231E and E160R/R237E led to constitutively open states of the channel [56]. Mutations in these critical S4 arginines alter the voltage dependence of channel activation. The effect of KCNE1 in these states is discussed in the later section. For example, R231A and R228E mutations causes the channel to become constitutively active/open [73,75]. Electrophysiology based patch-clamp experiments have shown that arginine R228 and R237 in S4 interact with E160, located towards the extracellular end of S2. R228 forms electrostatic interactions with E160 in the resting state whereas R237 pairs up with the same in the activated state of Q1-VSD, indicating the outward motion of S4 during Q1-VSD activation [74] (Figure 4A). An interesting feature of KCNQ1 is that its activation is both voltage dependent and mediated by a membrane lipid phosphatidylinositol 4,5-bisphosphate (PIP2). Depletion of PIP2 from the inner leaflet of the membrane causes suppression of KCNQ1 channel activity [75,76]. Zhang et al. studied the interaction of PIP2 with the open and closed states of KCNQ2 (Kv7.2) suggesting that PIP2 interacts with the S4–S5 linker region in the activated state and plays a role in gating mechanism, whereas in the closed state this region of contact lies in the S3–S3 loop of KCNQ2 [77]. In KCNQ1, basic residues in the S2–S3 and S4–S5 and the C-terminal region were identified as PIP2 binding sites with a higher PIP2 sensitivity in the presence of KCNE1 [77]. Voltage clamp fluorometry showed that PIP2 is required for the coupling of VSD-PD and in the presence of lipid phosphatase, CiVSP, the voltage sensor remains active/depolarized while the pore is closed [75,76,78]. PIP2 depletion causes disruption of the VSD-PD coupling and channel adopts an “uncoupled” state or is said to be “decoupled” [53,60]. The role of PIP2 as a coupling component is also supported by MD simulations [79]. In the resting state, PIP2 is located in the vicinity of S4 towards the cytosolic end of VSD. This prohibits the interaction of PIP2 with the amino acid residues on S6 helix of KCNQ1, thereby favoring a closed/resting state of the channel. In the active state, PIP2 forms salt bridges with the residues on S6 helix, shifting it further away from the positively charged residues on S4 favoring the active/open state of the channel [79]. Thus, PIP2 is suggested to favor the resting/closed and activated/open states of KCNQ1 [79]. The detailed structure of KCNQ1 in the decoupled state was recently published by Sun et al [53]. Based on the charge distribution and interaction of negatively charged PIP2 with the positively lined pockets between the voltage sensor and the pore forming region, two potential PIP2 binding regions were identified: One is the interface between Q1-VSD and PD (involving cytosolic end of S4 and S4–S5 linker), and the other region involves HB helix of the soluble C-terminal domain [60]. PIP2 has been shown to stabilize the open states of other Kv channels like hERG and KATP [80,81] and is thought to mediate similar effects in the KCNQ1-KCNE1 channel complex. Studies involving KCNQ1, KCNQ2, and hERG channels have shown that PIP2 migration towards the S2–S3 linker region regulates the deactivation kinetics of these channels [80,81]. The underlying mechanism of PIP2 migration in the case of hERG and KCNQ1/2 channels suggest that PIP2 interacts with the multiple binding sites on both these channels leading to acceleration of deactivation process [82].
4. The KCNQ1-KCNE1 Channel Complex
As previously stated, association of KCNQ1 with KCNE1 is crucial for the channel functioning. KCNE1 slows down the activation/deactivation kinetics of KCNQ1 and modulates the channel conductance critical for repolarization phase of cardiac action potential [31,63,83]. KCNE1 is a single transmembrane protein of 129 amino-acid residues with implications in channel biology and disease conditions [60]. Several studies have suggested that KCNE1 interacts with the PD of KCNQ1 [84–86]. Molecular modeling and studies involving the rat Kv1.2 channel structure showed unoccupied gaps between the VSDs, strengthening the claim that KCNE1 sits in the cleft between the adjacent VSDs and the PD [87,88] (Figure 5B). Cysteine crosslinking experiments involving the extracellular regions of KCNE1 and S1, S4 and S6 of KCNQ1 indicated state dependent interactions in these regions [83] (Figure 4B, 5B). Atrial Fibrillation (AF) related mutants in the S1 helix, S140G and V141M modulated the gating properties in the presence of KCNE1 [83]. Residues F57, T58, and L59 in the transmembrane helix of KCNE1 affected the voltage dependent channel modulation by interacting with residues S338, F339, F340 and A341 in the S6 region of KCNQ1 when mutated in either of the interacting partners [84,89,90]. These results suggest the involvement of S6 in KCNQ1-KCNE1 interaction and function. Cysteine accessibility experiments demonstrated a positive shift in the voltage dependence for T224C mutant in the presence of KCNE1 [91]. Mutations F232A in S4 and F279A in S5 helix showed changes in Q1-VSD movement and channel function in the presence of KCNE [92]. Similar experiments involving A226C in the S4 region slowed down the modification rate of methanethiosulfonate (MTS) reagent in the presence of KCNE1 indicating the stabilization of S4 in the closed state of Q1-VSD [93]. However, reports suggesting the association of KCNE1 altering the voltage dependence of the VSD and not the movement of S4 itself led to the conclusion that the observed changes affected the overall VSD equilibrium [94]. In the presence of KCNE1, the charge reversal mutations reported in the previous section showed interesting features. E160R/R231E currents were inhibited while the cell surface expression remained unaltered suggesting, 1) KCNE1 suppressed the intermediate-open state favoring an activated-open conformation, and 2) the above was a functional consequence on the intermediate-open state [56].
As we pointed out earlier, interaction with KCNE1 alters the overall topology of VSD-PD and has the potential to change the landscape of inter-subunit interaction as well. Based on these findings, it is obvious that KNCE1 modulates the overall functionality of KCNQ1 channel in several ways: by modulating VSD-PD coupling to prefer active-open state over the intermediate-open state, altering the ion selectivity and, increasing the current amplitude through enhanced VSD-PD association in the active-open state [56]. Despite the fragmented attempts to provide site specific interactions of KCNQ1 in the presence of KCNE1, voltage sensor movement and overall channel modulation, it is largely unclear as to how this 129 amino acid peptide regulates the Q1-VSD movement and modulates KCNQ1 activation kinetics on such a wide scale.
5. KCNQ1 and non-KCNE1 partners
While, the KCNQ1 activity modulation by KCNE1 has been studied to a great length, several other proteins (both membrane and cytosolic) and lipid molecules can directly affect the overall functionality of KCNQ1 [31,53,60,95]. One of these non-KCNE1 proteins is KCNE3. In an earlier study, it was suggested that in contrast to KCNE1, which affects both the VSD and gating of KCNQ1, KCNE3 only affects the VSD of the channel complex. The overall effect of KCNE3 is mediated through voltage sensor-to-gate coupling [96] (Figure 4C). Although, KCNE1 and KCNE3 display differences between their mode of action towards KCNQ1 (Figure 4B and 4C), their action is primarily controlled by the residues located in their TM segments. KCNQ1 forms a complex with KCNE3 and plays an important role in the recycling of potassium ions essential for transepithelial chloride ion secretion [97,98]. Transmembrane region of KCNE3 interacts with the open state of KCNQ1 and has been suggested to play an important role in the kinetics of initial binding and KCNQ1-KCNE3 complex formation. KCNE3 stabilizes the VSD region of KCNQ1 in its active state in a voltage dependent manner [99]. Electrophysiology based current measurements have shown that KCNE3 increases the KCNQ1-dependent current density by ~ 10-fold and leads to a constitutively active channel [99]. However, it is still not clear how KCNE3 affects the VSD but not gating of KCNQ1. Several studies have tried to explain the interaction mechanism between KCNQ1 and KNCE1 or, KCNE3 with the anticipation of discovering the interaction-function aspects. For example, Wrobel et al. have described the interactions between KCNQ1 and the N-terminal, the TM segment, and the C-terminal regions for both KCNE1 and KCNE3 (Figure 4B and 4C) [100]. Molecular modeling-based approach suggested that the extracellular end of the transmembrane region of KCNE3 sits in the cleft formed by extracellular end of S1 helix in one subunit and S5/P-loop/helix of the adjacent subunit of KCNQ1 [99]. Furthermore, it has also been proposed that either the pore region or the VSD of KCNQ1 channel could be directly interacting with the TM of KCNEs or, the C-terminus of the both of proteins may be interacting [65,66,84,85,93]. Even though we have several clues about the physical interaction regions/sites involved between KCNQ1 and KCNE1 or, KCNE3, it is still not clear how these interactions modulate the functionality of the overall channel complex. Therefore, this aspect of KCNQ1-KCNE3 biology still needs further exploration.
The activity of KCNQ1 is also modulated by polyunsaturated fatty acids (PUFA). Liin et al. showed that PUFA analogs act as IKS channel activators and play a role in cardiac exitability [101,102]. They proposed that PUFA and PUFA analogs incorporate their acyl chains into the lipid bilayer between the Q1-VSD, close to KCNE1 and bind to KCNQ1 via the lipid head group in the proximity of the positively charged amino acids of Q1-VSD [101–103]. The channel is activated as a result of this electrostatic interaction. The presence of KCNE1 leads to the protonation of PUFA, thereby exerting an opposite effect [103]. The negatively charged amino acid residues in the loop region connecting S5 to the pore helix play an important role in KCNE1 induced protonation [103] leading to the hypothesis that KCNE1 plays an indirect role in promoting PUFA protonation by inducing conformational changes in KCNQ1. These structural rearrangement result in the movement of the S5-p-helix loop closer to the PUFA binding site [103]. The PUFA analog, N-arachidonoyl taurine (N-AT) is shown to interact with the PD of KCNQ1 following electrostatic interaction with K326 in the S6 region [101]. N-AT increases the channel conductance by pulling K326 away from the pore region causing structural perturbations in the selectivity filter region of KCNQ1. PUFA analogs such as N-AT also interact with the positively charged S4 arginines in Q1-VSD and alter the voltage dependence by shifting it towards more negative voltages causing the channel to open [101].
As described earlier in this review, KCNQ1 plays various physiological roles in the heart as well as epithelial tissues. The multifunctional KCNQ1 is enabled by its interaction with several non-KCNEs proteins and cytosolic molecules acting as direct modulators of KCNQ1 functions. One of these modulators is protein kinase A (PKA). The β-adrenergic receptor stimulation plays an important role in KCNQ1-KCNE1 current generation during a high-stress, fast heart rate condition [104]. The β1-adrenergic receptor activation directly affects the intracellular level of cAMP leading to PKA activation, resulting in phosphorylation of the N-terminal region of KCNQ1. This leads to faster channel activation and, ultimately, shortening of the cardiac action potential. One of the proteins which plays important role in the anchoring of PKA to KCNQ1-KCNE1 complex is Yotiao or A-kinase anchoring protein (AKAP). It forms a macromolecular complex consisting of channel proteins, PKA, phosphodiesterase PDE4D3, protein phosphatase PP1, and adenylyl cyclase (AC9) in the heart [105,106,107]. Yotiao facilitates AC9- KCNQ1 complex formation in the heart following β-adrenergic stimulation [105,106,108]. Yotiao physically interacts with the residues in the cytosolic C-terminal domain of KCNQ1 and modulates the channel activity by affecting the phosphorylation/dephosphorylation events at S27 of the N-terminal end of KCNQ1 [106–108]. S27 phosphorylation does not require KCNE1. However, in the absence of KCNE1, this PKA-mediated phosphorylation of S27 does not result in hyperpolarization and channel activation, suggesting that KCNE1 plays a role in transducing the phosphorylation cascade to modulated channel function [109]. As described for membrane receptor proteins where phosphorylation plays an important role in activation/deactivation kinetics and disease manifestations [110,111], further understanding of the mechanism of PKA mediated post-translational modification of KCNQ1–KCNE1 activity will be important for understanding long QT mutations.
The cytosolic Ca2+-binding protein and Ca2+ signaling mediator, Calmodulin (CaM), plays important role in KCNQ1 channel assembly. CaM assists KCNQ1 in channel assembly to form functional tetramers [112,113]. In fact, when CaM does not interact with KCNQ1 due to mutations in the channel protein, it prevents tetramerization of KCNQ1, which is directly related with aberrant channel function and long QT syndrome. Furthermore, the cytosolic Ca2+ concentration affects the KCNQ1 activation. Physiologically reduced intracellular Ca2+ causes an inactivation of homomeric KCNQ1 [112]. CaM binds to the C-terminal of KCNQ1 and is essential for channel folding and assembly. Calmodulin binding results in a dimeric coiled-coil structure further forming a dimer of the dimer leading to tetramerization and a functional KCNQ1 [112,113]. The involvement of the C-terminal in KCNQ1 is unique because most of other Kv channels show this tetramerization mediated through their N-terminal region. The cryo-EM structure of the KCNQ1/CaM complex revealed two contact sites, 1) where the HA-HB helices of the soluble C-terminal segment of KCNQ1 insert in the middle of a clam-shell like structure formed by the N and C lobes of CaM, and, 2) between the S2–S3 loop of Q1-VSD (in the transmembrane region) and the third EF hand motif of CaM [53]. This second contact site along with the involvement of the fourth EF motif of CaM in the first contact site was not observed in the crystal structure reported earlier [114]. CaM interaction with the above two regions in KCNQ1 can be a useful link in unraveling the VGD-PD coupling mechanism. Recent studies by Chang et al. highlighted the role of an Apo/Cam clamp conformation of Cam C-lobe calcium dependent switch [115]. Apo/Cam clamp conformation, a common feature among all Kv7 isoforms, promotes the opening of Kv7.2-Kv7.5 channels but inhibits Kv7.1 (KCNQ1) activation. Binding of calcium to the C-lobe activates a CaM C-lobe causing its transition from semi-open to an open conformation, anchoring the N-lobe of CaM and perturbing the interactions with helix A leading to channel inhibition in Kv7.2- Kv7.5 while facilitating activation of Kv7.1 (KCNQ1) [115].
6. Conclusion and Future Directions
KCNQ1, one of the potassium ion channel family members, displays a multifunctional role and is important for various cellular processes/functions, for example cell volume homeostasis, electrical signaling in cardiac myocytes, glucose homeostasis, and gastric acid secretion. Aberrant KCNQ1 channel activity directly leads to disease phenotype and one of the prominent examples is long QT syndrome. KCNQ1 activity essentially requires its interaction with KCNE family member proteins, most prominently KCNE1. KCNQ1/KCNE1 interaction is further modulated by several other cellular proteins/molecules: calmodulin, PIP2, PKA, pH, and post translational modification (phosphorylation) of KCNQ1 itself. Due to these diverse interactions, KCNQ1 could perform several functions in the cellular physiology. As described in detail in this review, structural and biophysical studies of KCNQ1 have provided us several seminal findings showing the role of VSD and helix S5–S6 roles in the channel opening and gating. Even if we now know a significant amount regarding the roles for specific residues in KCNQ1 function and about its interactions with KCNE proteins, there are still many avenues of KCNQ1 structural-functional biology which needs to be explored in detail in future studies. Undoubtedly, future regulatory mechanistic studies of KCNQ1 function and its regulation will not only add to the knowledge base regarding KCNQ1 channel physiology under healthy and heart disease conditions, but will also illuminate its role in epithelial tissues, endocrinology, ear function and other cellular homeostasis functions.
Highlights:
KCNQ1 is related to several diseases, atrial fibrillation, long QT syndrome
KCNQ1 is ubiquitously expressed, complexes with members of KCNE family
Non-KCNE partners like PUFA, PIP2, CaM, PKA cause structure/functional changes
Acknowledgments
The authors thank the members of the Lorigan, Dabney-Smith, and Sanders research groups for their valuable suggestions throughout the development of this manuscript. This work was generously supported by the NIGMS/NIH Maximizing Investigator‟s Research Award (MIRA) R35 GM126935 award and a NSF CHE-1807131 grant (to Gary A Lorigan). The pulsed EPR spectrometer was purchased through funding provided by the NSF (MRI-1725502), the Ohio Board of Reagents, and Miami University. Gary A. Lorigan would also like to acknowledge support from the John W. Steube Professorship.
Abbreviations:
- LGIC
ligand-gated ion channel
- VGL
voltage-gated like
- nAchR
nicotinic acetylcholine receptor
- PIP2
phosphatidylinositol 4,5-bisphosphate
- IKS
slow delayed rectifier current
- PKA
protein kinase A
- PUFA
polyunsaturated fatty acid
- SNHL
sensorineural hearing loss
- AF
Atrial Fibrillation
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Conflict of interest
There is no conflict of interests.
Informed consent
The proposed article does not contain any unpublished studies.
References
- 1.Yu FH, Catterall WA. The VGL-chanome: A protein superfamily specialized for electrical signaling and ionic homeostasis. Sci STKE 2004; 5: re15. [DOI] [PubMed] [Google Scholar]
- 2.Hille B, Armstrong CM, MacKinnon R. Ion Channels: from idea to reality. Nat Med 1999; 5: 1105–1109. [DOI] [PubMed] [Google Scholar]
- 3.MacKinnon R. Potassium channels. FEBS Letters 2003; 555: 62–65. [DOI] [PubMed] [Google Scholar]
- 4.Lieve KV, Wilde AA. Inherited ion channel diseases: a brief review. Europace 2015; 17: 1–6. [DOI] [PubMed] [Google Scholar]
- 5.Ramakrishanan AM, Sankaranarayanan K. Understanding autoimmunity: the ion channel perspective. Autoimm Rev 2016; 15: 585–620. [DOI] [PubMed] [Google Scholar]
- 6.Bagal SK, Brown AD, Cox PJ, Omoto K, et al. Ion channels as therapeutic targets: a drug discovery perspective. J Med Chem 2013; 56: 593–624. [DOI] [PubMed] [Google Scholar]
- 7.Catterall WA. Structure and function of voltage-gated ion channels. Annual Review of Biochemstry 1995; 64: 493–531. [DOI] [PubMed] [Google Scholar]
- 8.Bezanilla F, Perozo E. The voltage sensor and gate in ion channels. Advances in Protein Chemistry 2003; 63: 211–241. [DOI] [PubMed] [Google Scholar]
- 9.Emmanouil DE, Quock RM. Advances in understanding the actions of nitrous oxide. Anesth Prog 2007; 54: 9–18. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 10.Neyroud N, Tesson F, Denjoy I, Leibovici M, Donger C, Barhanin J, Faure S, Gary F, Coumel P, Petit C, Schwartz K, Guicheney P. A novel mutation in the potassium channel gene KVLQT1 causes the Jervell and Lange-Nielsen cardioauditory syndrome. Nat Gen 1997; 15: 186–189. [DOI] [PubMed] [Google Scholar]
- 11.Tiron C, Campuzano O, Perez-Serra A, Mademont I, Coll M, Allegue C, Iglesias A, Partemi S, Striano P, Oliva A, Brugada R. Further evidence of the association between LQT syndrome and epilepsy in family with KCNQ1 pathogenic variant. Seizure 2015; 25: 65–67. [DOI] [PubMed] [Google Scholar]
- 12.Goldman AM, Glasscock E, Yoo J, Chen TT, Klassen TL, Noebels JL. Arrhythmia in heart and brain: KCNQ1 mutations link epilepsy and sudden unexplained death. Sci Transl Med 2009; 1: 2ra6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 13.Anantharam A, Markowitz SM, Abbott GW. Pharmacogenetic considerations in diseases of cardiac ion channels. J Pharmacol Exp Ther 2003; 307: 831–838. [DOI] [PubMed] [Google Scholar]
- 14.Lerche C, Seebohm G, Wagner CI, Scherer CR, Dehmelt L, Abitbol, Gerlach U, Brendel J, Attali B, and Busch AE. Molecular impact of MinK on the enantiospecific block of IKS by Chromanols. British Journal of Pharmacology 2000; 131: 1503–1506. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Veerman CC, Verkerk AO, Blom MT, Klemens CA, Langendijk Pim NJ, Van Ginneken ACG, Wilders R, and Tan HL. Slow delayed rectifier potassium current blockade contributes importantly to drug-induced Long-QT syndrome. Circ Arrhytm Electrophysiol 2013; 6(5): 1002–9. [DOI] [PubMed] [Google Scholar]
- 16.Maćias Á, Moreno C, Moral-Sanz J, Cogolludo Á, David M, Alemanni M, Pérez-Vizcaíno F, Zaza A, Valenzuela C, and González T. Celecoxib blocks cardiac Kv1.5, Kv4.3 and Kv7.1 (KCNQ1) channels effects on cardiac action potentials. Journal of Molecular and Cellular Cardiology 2010; 49: 984–992. [DOI] [PubMed] [Google Scholar]
- 17.Brueggemann LI, Mackie AR, Mani BK, Cribbs LL, and Byron KL. Differential effects of selective cyclooxygenase-2 inhibitors on vascular smooth muscle ion channels may account for differences in cardiovascular risk profiles. Mol Pharmacol 2009; 76: 1053–1061. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Smith JA, Vanoye CG, George AL, Meiler J, and Sanders CR. Structural models for the KCNQ1 voltage-gated potassium channel. Biochemistry 2007; 46, 14141–14152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Kapplinger JD, Tseng AS, Salisbury BA, Tester DJ, Callis TE, Alders M, Wilde AAM, and Ackerman MJ. Enhancing the predictive power of mutations in the C-terminus of the KCNQ1-encoded Kv7.1 voltage-gated potassium channel. J Cardiovasc Transl Res 2015; 8(3), 187–197. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Peng D, Kim JH, Kroncke BM, Law CL, Xia Y, Droege KD, Van Horn WD, Vanoye CG and Sanders CR. Purification and structural study of the voltage-sensor domain of the human KCNQ1 potassium ion channel. Biochemistry 2014; 53, 2032–2042. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 21.Chakrapani S, Cuello LG, Cortes MD, and Perozo E. Structural dynamics of an isolated-voltage sensor domain in lipid bilayer. Structure 2008; 16(3), 398–409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 22.Jiang Y, Lee A, Chen J, Ruta V, Cadene M, Chait BT, and MacKinnon R. X-Ray structure of a voltage-dependent K+ channel. Nature 2003; 423, 33–41. [DOI] [PubMed] [Google Scholar]
- 23.Ruta V, and Mackinnon R. Localization of the voltage-sensor toxin receptor on KvAP. Biochemistry 2004; 43, 10071–10079. [DOI] [PubMed] [Google Scholar]
- 24.Lee SY, Lee A, Chen J, MacKinnon R. Structure of the KvAP voltage-dependent K+ channel and its dependence on the lipid membrane. Proc Natl Acad Sci USA 2005; 102, 15441–15446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Long SB, Tao X, Campbell EB, MacKinnon R. Atomic structure of a voltage-dependent K+ channel in a lipid membrane-like environment. Nature 2007; 450, 376–382. [DOI] [PubMed] [Google Scholar]
- 26.Sahu ID, Craig AF, Dungan MM, Troxel K, Zhang R, Meiberg AG, Harmon CN, McCarrick RM, Kroncke BM, Sanders CR, Lorigan GA. Probing Structural dynamics and topology of the KCNE1 membrane protein in lipid bilayers via site-directed spin labeling and electron paramagnetic resonance spectroscopy. Biochemistry 2015; 5441, 6402–6412. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Barret PJ, Song Y, Van Horn WD, Hustedt EJ, Schafer JM, Hadziselimovic A, Beel AJ and Sanders CR. The amyloid precursor protein has a flexible transmembrane domain and binds cholesterol. Science 2012; 336, 1168–1171. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Sanders CR, Myers JK. Disease-related misassembly of membrane proteins. Annu Rev Biophys Biomol Struct 2004; 33: 25–51. [DOI] [PubMed] [Google Scholar]
- 29.Dwivedi P, Greis KD. Granulocyte colony-stimulating factor receptor signaling in severe congenital neutropenia, chronic neutrophilic leukemia, and related malignancies. Exp Hematol 2017; 46: 9–20. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Sanders CR, Nagy JK. Misfolding of membrane proteins in health and disease: the lady or the tiger? Curr Opin Struct Biol 2000; 10: 438–442. [DOI] [PubMed] [Google Scholar]
- 31.Abbott GW. Biology of the KCNQ1 potassium channel. New J Sci 2014; 237431. [Google Scholar]
- 32.Jespersen T, Grunnet M, Olsen SP. The KCNQ1 potassium channel: from gene to physiological function. Physiol 2005; 20: 408–416. [DOI] [PubMed] [Google Scholar]
- 33.Liin SI, Barro-Soria R, Larsson HP. The KCNQ1 channel-remarkable flexibility in gating allows for functional versatility. J Physiol 2015; 593: 2605–2615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Barhanin J, Lesage F, Guillemare E, Fink M, Lazdunski M, Romey G. KvLQT1 and IsK (mink) proteins associate to form the IKS cardiac potassium current. Nature 1996; 384: 78–80. [DOI] [PubMed] [Google Scholar]
- 35.Sanguinetti MC, Curran ME, Zou A, Shen J, Spector PS, Atkinson DL, Keating MT. Coassembly of KvLQT1 and minK (Isk) protein to form cardiac IKS potassium channel. Nature 1996; 384: 80–83. [DOI] [PubMed] [Google Scholar]
- 36.Wang KW, Tai KK, Goldstein SA. MinK residues line a potassium channel pore. Neuron 1996; 16: 571–577. [DOI] [PubMed] [Google Scholar]
- 37.Nerbonne JM, Kass RS. Molecular physiology of cardiac repolarization. Physiol Rev 2005; 85: 1205–1253. [DOI] [PubMed] [Google Scholar]
- 38.Hibino H, Nin F, Tsuzuki C, Kurachi Y. How is the highly positive endocochlear potential formed? The specific architecture of the stria vascularis and the roles of the ion-transport apparatus. Pflugers Arch 2010; 459: 521–533. [DOI] [PubMed] [Google Scholar]
- 39.Wangemann P. K+ recycling and the endocochlear potential. HearingResearch 2002; 165: 1–9. [DOI] [PubMed] [Google Scholar]
- 40.Hayashi M, Novak I. Molecular basis of potassium channels in pancreatic duct epithelial cells. Channels (Austin) 2013; 7: 432–441. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Warth R, Barhanin J. The multifaceted phenotype of the knockout mouse for the KCNE1 potassium channel gene. Am J Physiol Regul Integr Comp Physiol 2002; 282: R639–R648. [DOI] [PubMed] [Google Scholar]
- 42.Vallon V, Grahammer F, Volkl H, Sandu CD, Richter K, Rexhepaj R, Gerlach U, Rong Q, Pfeifer K, Lang F. KCNQ1-dependent transport in renal and gastrointestinal epithelia. Proc Natl Acad Sci USA 2005; 102: 17864–17869. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Schroeder BC, Waldegger S, Fehr S, Bleich M, Warth R, Greger R, Jentsch TJ. A constitutively open potassium channel formed by KCNQ1 and KCNE3. Nature 2000; 403: 196–199. [DOI] [PubMed] [Google Scholar]
- 44.Barrett KE, Keely SJ. Chloride secretion by the intestinal epithelium: molecular basis and regulatory aspects. Annu Rev Physiol 2000; 62: 535–572. [DOI] [PubMed] [Google Scholar]
- 45.Heitzmann D, Warth R. No potassium, no acid: K+ channels and gastric acid secretion. Physiology (Bethesda) 2007; 22: 335–341. [DOI] [PubMed] [Google Scholar]
- 46.Grahammer F, Warth R, Barhanin J, Bleich M, Hug MJ. The small conductance K+ channel, KCNQ1: expression, function, and subunit composition in murine trachea. J Biol Chem 2001; 276: 42268–42275. [DOI] [PubMed] [Google Scholar]
- 47.Preston P, Wartosch L, Gunzel D, Fromm M, Kongsuphol P, Ousingsawat J, Kunzelmann K, Barhanin J, Warth R, Jentsch TJ. Disruption of the K+ channel beta-subunit KCNE3 reveals an important role in intestinal and tracheal Cl− transport. J Biol Chem 2010; 285: 7165–7175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Roepke TK, Anantharam A, Kirchhoff P, Busque SM, Young JB, Geibel JP, Lerner DJ, Abbott GW. The KCNE2 potassium channel ancillary subunit is essential for gastric acid secretion. J Biol Chem 2006; 281: 23740–23747. [DOI] [PubMed] [Google Scholar]
- 49.Roepke TK, King EC, Reyna-Neyra A, Paroder M, Purtell K, Koba W, Fine E, Lerner DJ, Carrasco N, and Abbott GW. Kcne2 deletion uncovers its crucial role in thyroid hormone biosynthesis. Nat Med 2009; 15 (10): 1186–1194. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 50.Bartos DC, Anderson JB, Bastiaenen R, Johnson JN, Gollob MH, Tester DJ, Burgess DE, Homfray T, Behr ER, Ackerman MJ, Guicheney P, and Delisle BP. A KCNQ1 mutation causes a high penetrance for familial Atrial Fibrillation. J Cardiovas Electrophysiol 2013; 24 (5): 562–569. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Börjesson SI, Elinder F. Structure, function, and modification of the voltage sensor in voltage gated ion channels. Cell Biochem Biophys 2008; 52: 149–174. [DOI] [PubMed] [Google Scholar]
- 52.Catterall WA. Ion channel voltage sensors: structure, function, and pathophysiology. Neuron 2010; 67: 915–928. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 53.Sun J, and MacKinnon R. Cryo-EM structure of a KCNQ1/CaM complex reveals insights into congenital long QT syndrome. Cell 2017;169: 1042–1050. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Swartz KJ. Sensing voltage across lipid membranes. Nature 2008; 456: 891–897. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 55.del Camino D, Holmgren M, Liu Y, Yellen G. Blocker protection in the pore of a voltage-gated K+ channel and its structural implications. Nature 2000; 403: 321–325. [DOI] [PubMed] [Google Scholar]
- 56.Zaydman MA, Kasimova MA, McFarland K, Beller Z, Hou P, Kinser HE, Liang H, Zhang G, Shi J, Tarek M, Cui J. Domain–domain interactions determine the gating, permeation, pharmacology, and subunit modulation of the IKS ion channel. Elife 2014; 3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 57.Dwivedi P, Rodriguez J, Ibe NU, Weers PM. Deletion of the N- and C-terminal helix of apolipophorin III to create a four-helix bundle protein. Biochemistry 2016; 55 (26): 3607–3615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 58.Bellesia G, Jewett AI, Shea JE. Relative stability of de novo four-helix bundle proteins: insights from corse grained molecular simulations. Protein Sci 2011; 20: 818–826. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 59.Kohn WD, Mant CT, Hodges RS. α-helical protein assembly motifs. J Biol Chem 1997; 272: 2583–2586. [DOI] [PubMed] [Google Scholar]
- 60.Cui J. Voltage-dependent gating: Novel insights from KCNQ1 channels. Biophysical Journal 2016; 110: 14–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Tao X, Lee A, Limapichat W, Dougherty DA, MacKinnon R. A gating charge transfer center in voltage sensors. Science 2010; 328: 67–73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 62.Jiang Y, Lee A, Chen J, Cadene M, Chait BT, MacKinnon R. The open pore conformation of potassium channels. Nature 2002; 417: 523–526. [DOI] [PubMed] [Google Scholar]
- 63.Nakajo K, Ulbrich MH, Kubo Y, Isacoff EY. Stoichiometry of the KCNQ1–KCNE1 ion channel complex. Proc Natl Acad Sci USA 2010; 107: 18862–18867. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Plant LD, Xiong D, Dai H, Goldstein SA. Individual IKS channels at the surface of mammalian cells contain two KCNE1 accessory subunits. Proc Natl Acad Sci USA 2014; 111: E1438–1446. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 65.Xu X, Jiang M, Hsu KL, Zhang M, Tseng GN. KCNQ1 and KCNE1 in the IKS channel complex make state dependent contacts in their extracellular domains. J Gen Physiol 2008; 131: 589–603. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 66.Chung DY, Chan PJ, Bankston JR, Yang L, Liu G, Marx SO, Karlin A, Kass RS (2009). Location of KCNE1 relative to KCNQ1 in the IKS potassium channel by disulfide cross-linking of substituted cysteines. Proc Natl Acad Sci USA 2009; 106: 743–748. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 67.Sanguinetti MC, Curran ME, Zou A, Shen J, Specter PS, Atkinson DL, and Keating MT. Coassembly of KvLQT1 and mink (IsK) proteins to form cardiac IKS potassium channel. Nature 1996; 384: 80–83. [DOI] [PubMed] [Google Scholar]
- 68.Hasani HJ, Ganesan A, Ahmed M, and Barakat KH. Effects of protein-protein interactions and ligand binding on the ion permeation in KCNQ1 potassium channel. PLoS One 2018; 13 (2): e0191905. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Xu Y, Wang Y, Meng XY, Zhang M, Jiang M, Cui M, and Tseng GN. Building KCNQ1/KCNE1 channel models and probing their interactions by molecular dynamic simulations. Biophysical Journal 2013; 105: 2461–2473. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 70.Wu D, Pan H, Cui J. 2010. KCNE1 remodels the voltage sensor of Kv7.1 to modulate channel function. Biophys J 2010; 99:3599–3608. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Papazian DM, Shao XM, Seoh SA, Mock AF, Huang Y, Wainstock DH. Electrostatic interactions of S4 voltage sensor in Shaker K+ channel.Neuron 1995; 14: 1293–1301. [DOI] [PubMed] [Google Scholar]
- 72.Dixit G, Sahu ID, Reynolds WD, Wadsworth TM, Harding BD, Jaycox CK, Dabney-Smith C, Sanders CR, Lorigan GA. Probing the dynamics and structural topology of the reconstituted human KCNQ1 voltage sensor domain (Q1-VSD) in lipid bilayers using electron paramagnetic resonance spectroscopy. Biochemistry 2019; 58: 965–973. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 73.Aggarwal SK, and MacKinnon R. Contribution of the S4 segment to gating charge in the shaker K+ channel. Neuron, 1996; 16: 1169–1177. [DOI] [PubMed] [Google Scholar]
- 74.Wu D, Delaloye K, Zaydman MA, Nekouzadeh A, Rudy Y, and Cui J. State-dependent electrostatic interactions of S4 arginines with E1 in S2 during Kv7.1 activation. J Gen Physiol, 2010; 135: 595–606. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 75.Zaydman MA, Silva JR, Delaloye K, Li y, Liang H, Larson HP, Shi J, and Cui J. Kv7.1 ion channels require a lipid to couple voltage sensing to pore opening. Proc. Natl. Acad. Sci. USA 2013; 110: 13180–13185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Suh BC, and Hille B. Regulation of ion channels by phosphatidyl inositol-4,5-bisphosphate. Curre. Opin. Neurobiol 2005; 15: 370–378. [DOI] [PubMed] [Google Scholar]
- 77.Zhang Q, Zhou P, Chen Z, Li M, Jiang H, Gao Z, and Yang H. Dynamic PIP2 interactions with voltage sensor elements contribute to KCNQ2 channel gating. PNAS 2013; 110 (50): 20093–20098. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Zaydman MA, and Cui J. PIP2 regulation of KCNQ channels: biophysical and molecular mechanisms for lipid modulation of voltage-dependent gating. Front. Physiol 2014; 5:195. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 79.Kasimova MA, Zaydman MA, Cui J, and Tarek M. PIP2 dependent coupling is prominent in Kv7.1 due to weakened interactions between S4–S5 and S6. Scientific Reports 2014; 5: 7474. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Rodriguez N, Amarouch MY, Montnach J, Piron J, Labro AJ, Charpentier F, Merot J, Baro I, and Loussouarn G. Phosphatidyl inositol-4,5-bisphosphate (PIP2) stabilizes the open pore conformation of of the Kv11.1 (hERG) channel. Biophysical Journal 2010; 99: 1110–1118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 81.Ribalet B, John SA, and Weiss JN. Regulation of cloned ATP-sensitive K channels by phosphorylation,MgADP, and phosphatidyl inositol-4,5-bisphosphate (PIP2): A study of channel rundown and reactivation. Journal of General Physiology 2000; 116: 391–409. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Chen L, Zhang Q, Qiu Y, Li Z, Chen Z, Jiang H, Li Y, and Yang H. Migration of PIP2 on voltage gated potassium channel surface influences channel deactivation. Scientific Reports 2015; 5: 15079. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 83.Nakajo K, and Kubo Y. KCNQ1 channel modulation by KCNE proteins via the voltage sensing domain. J. Physiol 2015; 593: 2617–25. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Panaghie GK, Tai KK, and Abbott GW. Interaction of KCNE subunits with the KCNQ1 K+ channel pore. J. Physiol 2006; 570: 455–467. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Melman YF, Um SY, Krumerman A, Kagan A, McDonald TV. KCNE1 binds to the KCNQ1 pore to regulate potassium channel activity. Neuron. 2004; 42: 927–937. [DOI] [PubMed] [Google Scholar]
- 86.Panaghie G, Abbott GW. The role of S4 charges in voltage-dependent and voltage-independent KCNQ1 potassium channel complexes. J Gen Physiol 2007; 129:121–133. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 87.Van Horn WD, Vanoye CG, Sanders CR. Working model for the structural basis for KCNE1 modulation of the KCNQ1 potassium channel. Curr. Opin. Struct. Biol 2011; 21: 283–291. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Kang C, Tian C, Sonnichsen FD, Smith JA, Meiler J, George AL, Vanoye CG, Kim HJ, Sanders CR. Structure of KCNE1 and implications for how it modulates the KCNQ1 potassium ion channel. Biochemistry 2008; 47: 7999–8006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 89.Melman YF, Domenech A, de la LS, and McDonald TV. Structure determinants of KvLQT1 control by KCNE family of proteins. J. Biol. Chem 2001; 276: 6439–6444. [DOI] [PubMed] [Google Scholar]
- 90.Mikuni I, Torres CG, Bienengraeber MW, and Kwok WM. Partial restoration of the long QT syndrome associated KCNQ1 A341V mutant by the KCNE1 β-subunit. Biochim. Biophys. Acta 2011; 1810: 1285–1293. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 91.Barro-Soria R, Rebolledo S, Liin SI, Perez ME, Sampson KJ, Kass RS, and Larsson HP. KCNE1 divides the voltage sensor movement in KCNQ1/KCNE1 channels into two steps. Nat. Commun 2014; 5: 3750. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 92.Nakajo K, and Kubo Y. Steric hindrance between S4 and S5 of the KCNQ1/KCNE1 channel hampers pore opening. Nat. Commun 2014; 5: 4100. [DOI] [PubMed] [Google Scholar]
- 93.Nakajo K, and Kubo Y. KCNE1 and KCNE3 stabilize and/ or slow voltage sensing S4 segment of KCNQ1 channel. J Gen. Physiol 2007; 130: 269–281. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Rocheleau JM, and Kobertz WR. KCNE peptides differently affect voltage sensor equilibrium and equilibration rates in KCNQ1 K+ channels. J. Gen. Phyisiol 2008; 131: 59–68. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 95.Taylor KC, Sanders CR. Regulation of KCNQ/Kv7 family voltage-gated K+ channels by lipids. Biochim Biophys Acta 2017; 1859: 586–597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 96.Barro-Soria R, Perez ME, and Larsson HP. KCNE3 acts by promoting voltage sensor activation in KCNQ1. Proc. Natl. Acad. Sci USA 2015; 112: E7286–E7292. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 97.Abbott GW. KCNE1 and KCNE3: The yin and yang of voltage-gated K+ channel regulation. Gene 2016; 576: 1–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 98.Preston P, Wartosch L, Gunzel D, Fromm M, Kongsuphol P, Ousingsawat J, Kunzelmann K, Barhanin J, Warth R, and Jentsch TJ. Disruption of the K+ channel β-subunit KCNE3 reveals an important role in intestinal and tracheal Cl− transport. J Biol Chem. 2010; 285: 7165–7175. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 99.Kroncke BM, Van Horn WD, Smith J, Kang CB, Welch RC, Song Y, Nannemann DP, Taylor KC, Sisco NJ, George AL, Meiler J, Vanoye CG, and Sanders CR. Structural basis for KCNE3 modulation of potassium recycling in epithelia. Sci Adv 2016; 2: e1501228. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 100.Wrobel E, Tapken D, and Seebohm G. The KCNE tango-how KCNE1 interacts with Kv7.1. Front Pharmacol 2012; 3: 142. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 101.Liin SI, Yazdi S, Ramentol R, Barro-Soria R, and Larsson PH. Mechanisms underlying the dual effect of polyunsaturated fatty acid analogs on Kv7.1. Cell Reports 2018; 24: 2908–2918. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Liin SI, Ejneby MS, Barro-Soria R, Skarsfekdt MA, Larsson JE, Härlin FS, Parkkari T, Bentzen BH, Schmitt N, Larsson PH, and Elinder F. Polyunsaturated fatty acid analogs act antiarrhythmically on the cardiac IKS channel. PNAS 2015; 112: 5714–5719. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Larsson JE, Larsson PH, and Liin SI. KCNE1 tunes the sensitivity of Kv7.1 to Polyunsaturated fatty acids by moving turret residues close to the binding site. eLIFE 2018; 7: e37257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Terrenoire C, Clancy CE, Cormier JW, Sampson KJ, and Kass RS. Autonomic control of cardiac action potentials: role of potassium channel kinetics in response to sympathetic stimulation. Ciec Res 2005; 96: e25–e34. [DOI] [PubMed] [Google Scholar]
- 105.Marx SO, Kurokawa J, Reiken S, Motoike H, D‟Armiento J, Marks AR, and Kass RS. Requirement of a macromolecular signaling complex for beta adrenergic receptor modulation of the KCNQ1-KCNE1 potassium channel. Science 2002; 295: 496–499. [DOI] [PubMed] [Google Scholar]
- 106.Kurokawa J, Motoike HK, Rao J, and Kass RS. Regulatory actions of the A-kinase anchoring protein Yotiao on a heart potassium channel downstream of PKA phosphorylation. Prot Natl Acad Sci USA 2004; 101: 16374–16378. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Li Y, Chen L, Kass RS, and Dessauer CW. The A-kinase anchoring protein Yotiao facilitates complex formation between adenylyl cyclase type 9 and the IKS potassium channel in heart. J Biol Chem 2012; 287 (35): 29815–24. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Dvir M, Peretz A, Haitin Y, and Attali B. Recent molecular insights from mutated IKS channels in cardiac arrhythmia. Curr Opin Pharmacol 2014; 15: 74–82. [DOI] [PubMed] [Google Scholar]
- 109.Kurokawa J, Chen L, and Kass RS. Requirement of subunit expression for cAMP-mediated regulation of a heart potassium channel. PNAS 2003; 100 (4): 2122–2127. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 110.Dwivedi P, Muench DE, Wagner M, Azam M, Grimes HL, Greis KD. Time resolved quantitative phospho-tyrosine analysis reveals Bruton‟s tyrosine kinase mediated signaling downstream of the mutated granulocyte-colony stimulating factor receptors. Leukemia 2019; 33(1): 75–87. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 111.Dwivedi P, Muench DE, Wagner M, Azam M, Grimes HL, Greis KD. Phospho serine and threonine analysis of normal and mutated granulocyte colony stimulating factor receptors. Sci Data 2019; 6(1): 21. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Ghosh S, Nunziato DA, and Pitt GS. KCNQ1 assembly and function is blocked by long-QT syndrome mutations that disrupt interaction with calmodulin. Circulation Research 2006; 98: 1048–1054. [DOI] [PubMed] [Google Scholar]
- 113.Shamgar L, Ma L, Schmitt N, Haitin Y, Peretz A, Wiener R, Hirsch J, Pongs O, and Attali B. Calmodulin is essential for cardiac IKS channel gating and assembly: impared function in long-QT mutations. Circulation Research 2006; 98: 1055–1063. [DOI] [PubMed] [Google Scholar]
- 114.Sachyani D, Dvir M, Strulovich R, Tria G, Tobelaim W, Peretz A, Pongs O, Svergun D, Attali B, and Hirsch JA. Structural basis of a Kv7.1 potassium channel gating module: studies of the intracellular c-terminal domain in complex with calmodulin. Structure 2014; 22: 1582–1594. [DOI] [PubMed] [Google Scholar]
- 115.Chang A, Abderemane-Ali F, Hura GL, Rossen ND, Gate RE, and Minor DL Jr. A calmodulin C-lobe Ca2+- dependent switch governs Kv7 channel function. Neuron 2018; 97: 836–852. [DOI] [PMC free article] [PubMed] [Google Scholar]





