Next Article in Journal
Hydroxycinnamic Acid Derivatives from Coffee Extracts Prevent Amyloid Transformation of Alpha-Synuclein
Next Article in Special Issue
Depletion of CD4 and CD8 T Cells Reduces Acute Disease and Is Not Associated with Hearing Loss in ML29-Infected STAT1-/- Mice
Previous Article in Journal
Irisin Is Related to Non-Alcoholic Fatty Liver Disease (NAFLD)
Previous Article in Special Issue
Two Loci Contribute to Age-Related Hearing Loss Resistance in the Japanese Wild-Derived Inbred MSM/Ms Mice
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Pathological Mechanisms of Hearing Loss Caused by KCNQ1 and KCNQ4 Variants

1
Department of Otolaryngology-Head and Neck Surgery, Feinberg School of Medicine, Northwestern University, Chicago, IL 60611, USA
2
The Hugh Knowles Center for Clinical and Basic Science in Hearing and Its Disorders, Northwestern University, Evanston, IL 60608, USA
Biomedicines 2022, 10(9), 2254; https://doi.org/10.3390/biomedicines10092254
Submission received: 5 August 2022 / Revised: 8 September 2022 / Accepted: 9 September 2022 / Published: 12 September 2022
(This article belongs to the Special Issue Genetic Research on Hearing Loss)

Abstract

:
Deafness-associated genes KCNQ1 (also associated with heart diseases) and KCNQ4 (only associated with hearing loss) encode the homotetrameric voltage-gated potassium ion channels Kv7.1 and Kv7.4, respectively. To date, over 700 KCNQ1 and over 70 KCNQ4 variants have been identified in patients. The vast majority of these variants are inherited dominantly, and their pathogenicity is often explained by dominant-negative inhibition or haploinsufficiency. Our recent study unexpectedly identified cell-death-inducing cytotoxicity in several Kv7.1 and Kv7.4 variants. Elucidation of this cytotoxicity mechanism and identification of its modifiers (drugs) have great potential for aiding the development of a novel pharmacological strategy against many pathogenic KCNQ variants. The purpose of this review is to disseminate this emerging pathological role of Kv7 variants and to underscore the importance of experimentally characterizing disease-associated variants.

1. Introduction

The human genome encodes 40 voltage-dependent potassium ion (K+) channels that belong to 12 subfamilies (Kv1–Kv12). The voltage-dependent activities of Kv channels play critical roles in controlling the electrophysiological properties of cells and maintaining the ion homeostasis. All Kv channels possess six transmembrane (TM) segments (S1–S6). The first four TM segments (S1–S4) constitute the voltage-sensing domain, while the last two, which flank a channel pore loop (P-loop), constitute the pore-forming domain (S5-PH-S6) (Figure 1). Tetramerization is essential for completing the K+-selective channel pore in all Kv channel types. Two Kv channels, Kv7.1 and Kv7.4, are essential for normal operation of the inner ear [1,2].
Kv7.1 is encoded by KCNQ1. It is abundantly expressed in the heart and is crucial for normal repolarization of cardiomyocytes. Mutations in Kv7.1 underlie two forms of the long QT syndrome (LQTS), i.e., the Romano–Ward syndrome (RWS) [3] and Jervell and Lange–Nielsen syndrome (JLNS) [4]. Cardiac symptoms in JLNS are typically more severe than those in RWS, and most patients with JLNS also suffer from congenital hearing loss [4]. RWS is inherited dominantly, with some exceptions [5,6], whereas JLNS is inherited recessively. In the cochlea, Kv7.1 is expressed in the marginal cells of the stria vascularis (SV) [4]. Kv7.1 is thought to mediate the secretion of K+ into the endolymph and the establishment of the endocochlear potential (EP) [7].
Kv7.4 is encoded by KCNQ4. In the cochlea, two types of sound-sensing cells, inner hair cells (IHCs) and outer hair cells (OHCs), are housed in the organ of Corti, and their apical surfaces exposed to the endolymph. Kv7.4 is abundantly expressed in OHCs, but it is also expressed in IHCs and the spiral ganglion neurons (SGNs) [8,9]. The K+ conductance mediated by Kv7.4 contributes to the establishment of a normal resting membrane potential and is crucial for repolarization of the cells after sound-elicited cell depolarization. The large Kv7.4-mediated conductance in OHCs also contributes to the reduction of the membrane time constant so that the receptor-potential-induced mechanical response of OHCs, i.e., electromotility [10,11], can respond to sound stimuli at high frequencies [12]. K+ that flows into OHCs from the high-K+-containing endolymph via the stereocilia is thought to be extruded by Kv7.4 from the base of OHCs to the perilymph. This Kv7.4-mediated extrusion of K+ is believed to be crucial for maintaining the intracellular ionic homeostasis and, thus, for OHC maintenance. Mutations in Kv7.4 are responsible for dominantly inherited progressive nonsyndromic hearing loss, DFNA2A [13].

2. The Pathologies of JLNS and DFNA2A Hearing Loss

Hearing loss in JLNS patients is congenital, bilateral, and profound (OMIN: 220400). To date, 768 KCNQ1 variants have been reported, ~5% of which are JLNS-associated (the Human Gene Mutation Database, HGMD) [14] (Table 1). The ionic composition of endolymph is unique among extracellular fluids in that it contains high K+ (~160 mM) but low Na+ (~1 mM) and Ca2+ (~20 µM) [15]. The high K+ in endolymph is due to the secretion of K+ from the marginal cells of the SV, which is powered by ATP-dependent Na+/K+ pumps, a Na+/K+/2Cl cotransporter, and other ion transporters. K+ transport from the marginal cells to the endolymph is electrogenic, resulting in positive EP (+80 to +100 mV) [16]. This very positive EP is a main driving force for sensory mechanotransduction by both hair cells (conversion of sound-induced mechanical vibrations sensed by stereocilia into changes in the receptor potential). Kv7.1, together with its ancillary protein, KCNE1, is thought to mediate K+ transport from the marginal cells to the endolymph. Thus, JLNS-causing KCNQ1 variants are presumed to affect EP by impairing the K+ conductance of Kv7.1. In fact, Kcnq1/ mice lose EP and show atrophy of the SV, reduction of the endolymphatic compartment, and degeneration of the organ of Corti and SGNs [17,18,19,20].
Hearing loss in DFNA2A patients is typically progressive and more prominent at higher frequencies (middle and low frequencies are also affected later in life) (OMIN: 600101). The severity of DFNA2A hearing loss and the rate of progression vary among KCNQ4 variants. To date, 76 KCNQ4 variants have been reported in HGMD (Table 2). Given the multiple aforementioned roles of Kv7.4 in OHCs, it is anticipated that DFNA2A-causing KCNQ4 variants impair normal cochlear operation by primarily affecting OHCs. Kcnq4/, Kcnq4G286S/+, and Kcnq4G286S/G286S (p.G285S in humans) mice recapitulate progressive hearing loss that is, indeed, accompanied by OHC dysfunction and degeneration [52]. Degeneration of IHCs and SGNs at later postnatal ages was also found [53].
The presence of multiple JLNS-/DFNA2A-associated variants found in KCNQ1/KCNQ4 and identification of hearing phenotypes in Kcnq1 and Kcnq4 mouse models compellingly establish the essentiality of these Kv7 channels in hearing.

3. The Pathological Roles of Kv7.1 and Kv7.4 Variants and Ongoing Pharmacological Strategies

A dominant-negative inhibition (inhibition of the Kv7 channel function or membrane targeting by having a mutated subunit in the tetramer complex) often underlies the observed dominant inheritance of Kv7.1 and Kv7.4 variants associated with RWS and DFNA2A. Since tetramerization of Kv7 channels is mediated by the C-terminal cytosolic helices C and D (HC and HD in Figure 1) [95,96,97,98,99,100,101,102,103], haploinsufficiency is presumed to account for the dominant inheritance of frame-shifted or nonsense Kv7 variants that lack the C-terminal tetramerization region. In any case, the functional consequence of Kv7 variants would be the reduction of overall K+ channel activity. Hence, pharmacological augmentation of reduced residual Kv7 channel activity by channel openers is considered clinically effective and actively pursued [54,102,103,104,105,106,107,108,109,110,111,112,113,114].

4. Observations against a Haploinsufficiency-Based Pathological Mechanism

Kv7.1 variants associated with JLNS are inherited recessively. In other words, individuals heterozygous for JLNS-associated Kv7.1 variants suffer from neither LQTS nor hearing loss. This suggests that one functional KCNQ1 allele is sufficient for maintaining normal cardiac and auditory functions, thus arguing against a haploinsufficiency-based pathological mechanism. It is perplexing that, among truncated Kv7.1 variants lacking the C-terminal tetramerization region, some are inherited dominantly, while the others are inherited recessively.
Multiple variants truncating the C-terminal tetramerization region are also identified in KCNQ4. If haploinsufficiency accounted for dominant inheritance of these truncated Kv7.4 variants, they should all result in a similar and a relatively mild DFNA2A phenotype compared to those that exert a dominant-negative inhibitory effect on the wild-type Kv7.4 (Kv7.4WT) subunit. The slowly progressive hearing loss found in patients with heterozygous Kv7.4Q71Sfs (c.211delC) [90] is in line with such a view. However, Kv7.4W242X (c.725G>A), which completely lacks the channel pore-forming transmembrane domain and the following C-terminal tetramerization region (Figure 1), is associated with severe to profound hearing loss [64]. The presence of recessively inherited truncated Kv7.4 variant, Kv7.4A349Pfs (c.1044_1051del8) [92], is also incompatible with a haploinsufficiency-based pathological mechanism. Incidentally, Kcnq4+/ heterozygous mice do not suffer from hearing loss [52], suggesting that one functional KCNQ4 allele is sufficient for maintaining normal auditory function, at least in mice.

5. Identification of Cell Death-Inducing Cytotoxicity in Truncated Kv7.1 and Kv7.4 Variants

In order to elucidate the pathogenic roles of truncated Kv7 variants, we first tried to confirm the absence of K+ channel activity for three deafness-associated Kv7.4 variants that completely lack the C-terminal tetramerization region, Kv7.4Q71Sfs, Kv7.4W242X, and Kv7.4A349Pfs (Figure 1), in HEK293T-based stable cell lines. Unexpectedly, we encountered great difficulties in establishing stable cell lines that were to constitutively express these variants, especially for Kv7.4W242X. After confirming the reproducibility of this issue, we suspected inherent cytotoxicity in these truncated variants. By performing CellTox Green Cytotoxicity and RealTime-Glo Annexin V Apoptosis assays using doxycycline-inducible stable cell lines, we found that these three Kv7.4 variants are indeed cytotoxic and induce cell death to various degrees in a doxycycline-dosage-dependent manner [41]. The cell-death-inducing cytotoxicity of these Kv7.4 variants was further confirmed in an HEI-OC1 cell line that was derived from the murine inner ear [41]. We also found that none of these truncated Kv7.4 variants were functional by themselves (when singly expressed); nor were they capable of forming a heteromer with Kv7.4WT [41], refuting the possibility that Kv7.4W242X and Kv7.4A349Pfs physically interact with Kv7.4WT to exert either an inhibitory (dominant-negative) or cooperative effect.
A CellTox Green Cytotoxicity assay was also performed for HEK293T-based doxycycline-inducible stable cell lines expressing truncated Kv7.1 variants lacking the C-terminal tetramerization domain, Kv7.1E261X, Kv7.1W305X, Kv7.1Q530X, and Kv7.1Q531X (Figure 1), among which solely Kv7.1Q530X was JLNS-associated [115]. Large cell-death-inducing cytotoxicity was found in Kv7.1E261X and Kv7.1W305X, while small cytotoxicity was found in Kv7.1Q530X and Kv7.1Q531X. A previous study showed that Kv7.1Q530X does not exert a dominant-negative inhibitory effect on wild-type Kv7.1 (Kv7.1WT) [5,29]. Consistently, another study showed that Kv7.1Q530X does not bind to Kv7.1WT [31]. It is probable that the other three Kv7.1 variants, which are truncated similarly (Kv7.1Q531X) or are shorter (Kv7.1E261X and Kv7.1W305X) compared to Kv7.1Q530X, are also incapable of physically interacting with Kv7.1WT.
Collectively, these observations affirm the inadequacy of a pathological mechanism that is based on dominant-negative inhibition or haploinsufficiency alone for explaining the mode of inheritance of some Kv7 variants.

6. The Mechanism Underlying Cell-Death-Inducing Cytotoxicity

Both Kv7.1 and Kv7.4 are membrane proteins that are translated and matured in the ER. It is conceivable that misfolding of truncated Kv7 variants and their accumulation induce chronic cellular stress and eventual cell death. In fact, we detected splicing of XBP1 mRNA, which is indicative of ER stress, in cells expressing Kv7.4 variants [41]. We also found that application of autophagy inducers, imatinib, SB202190, and FK-506, delayed the onset of cell death induced by the three truncated Kv7.4 variants [41], suggesting that accumulation of the truncated Kv7.4 proteins indeed underlies cell-death-inducing cytotoxicity. It should be noted that XBP1 mRNA splicing and small but nonnegligible cell-death-inducing cytotoxicity were also detected in cells expressing Kv7.4WT [41]. The small cytotoxicity of Kv7.4WT may be of physiological relevance because Kv7.4 is abundantly expressed in OHCs and because doxycycline-driven forced overexpression of prestin, another membrane protein abundantly expressed in OHCs, induced much less cell death compared to Kv7.4WT [41]. It may be conceivable that the Kv7.4 protein tends to misfold and self-aggregate even for the wild type, and that the basal endogenous autophagy activity needs to be kept high in OHCs to counteract the cytotoxicity associated with the large expression of Kv7.4. Such a view is in line with a recent study demonstrating that an inhibitor of autophagy, chloroquine, specifically damages OHCs in mice [116]. Pharmacological intervention to maintain or augment the endogenous autophagy activity may delay or prevent age-related hearing loss (presbycusis).

7. Updates on the Pathological Mechanisms of JLNS and DFNA2A

A dominant-negative inhibition-based pathological mechanism has been experimentally validated for many dominantly inherited Kv7.1 and Kv7.4 variants. We confirmed that two missense Kv7.4 variants, Kv7.4G285S and Kv7.4P291L, can interact with Kv7.4WT, and that their cytotoxicity is WT-like [41]. Thus, our study does not challenge the dominant-negative inhibition-based pathological mechanism (at least for these two Kv7.4 missense variants). However, to the best of my knowledge, a haploinsufficiency-based pathological mechanism has never been firmly demonstrated for any Kv7 variants. As mentioned above, a haploinsufficiency-based pathological mechanism is incompatible with several observations and, thus, not compelling.
The cell-death-inducing cytotoxicity identified in our recent study [41] provides a straightforward explanation for the observed dominant inheritance of truncated Kv7 variants lacking the ability to exert a dominant-negative inhibitory effect. It can also explain why some Kv7.1 variants that are associated with severe cardiac phenotypes are not always co-associated with proportionally longer corrected QT (QTc) intervals [117,118,119,120]. For example, a sixteen-year-old proband with Kv7.1W305X died suddenly, and two other affected members experienced syncopal episodes, although the resting QTc interval of patients with this Kv7.1 variant was found to be normal to borderline [120]. The normal-like QTc interval seems reasonable because it is unlikely that Kv7.1W305X exerts a dominant-negative inhibitory effect on Kv7.1WT and because one functional KCNQ1 allele is sufficient for supporting normal cardiac function (see above). It is conceivable that the severe cardiac phenotypes found in patients with Kv7.1W305X are ascribed to the large cell-death-inducing cytotoxicity found in this variant [41]. The Kv7.1WT-like small cytotoxicity found in recessively inherited Kv7.1Q530X among four truncated Kv7.1 variants tested [41] is compatible with such a view.
The vast majority of Kv7.1 variants are inherited dominantly and are associated with LQTS (RWS), but not with hearing loss. Homozygous or compound heterozygous Kv7.1 mutations that are not associated with hearing loss are also reported [5,6]. These observations suggest that the cardiac function is more sensitive to the reduction of Kv7.1 function compared to the auditory function. Thus, as it has been presumed, JLNS hearing loss is probably caused by drastic or complete loss of Kv7.1-mediated K+ channel activity. Pathological contributions of cytotoxicity of Kv7.1 variants are likely minimal in JLNS, given the observed recessive inheritance of JLNS variants.

8. Remaining Questions and Future Directions

It seems reasonable to speculate that the severity of disease phenotypes and the mode of inheritance of truncated Kv7 variants lacking their C-terminal tetramerization regions is largely determined by the magnitude of their anticipated cytotoxicity. However, the large cytotoxicity found in Kv7.4A349Pfs contradicts the fact that this variant is inherited recessively [92]. Future studies should address if the truncated Kv7 variants are actually expressed in natural host cells. In addition, the vulnerability of natural host cells to cytotoxic Kv7 variants may be very different from those of HEK293T and other cell lines. Animal models need to be generated to fully examine the pathophysiological role of cytotoxic Kv7 variants found in our in vitro study. Animal models are also needed to examine potential adverse effects of drugs (e.g., autophagy inducers). For example, imatinib was found to delay the onset of cell death induced by cytotoxic truncated Kv7.4 variants [41]; however, it was reported to be ototoxic [121,122]. We are currently in the process of generating knock-in mouse models for several truncating Kcnq4 variants to address these questions.
Despite its limitations, in vitro characterization of disease-associated Kv7 variants using doxycycline-inducible stable cell lines, as showcased in our recent study for Kv7.1 and Kv7.4 [41], will remain useful as an experimental platform for rapidly identifying cytotoxic variants and screening for potentially effective drugs. In general, a dominant-negative inhibitory effect is suspected for dominantly inherited missense Kv7 variants that retain the C-terminal tetramerization region. For example, for Kv7.4 missense variants, a dominant-negative inhibitory effect has been demonstrated for p.L47P, p.N264S, p.S269F, p.S273A, p.L274H, p.W276S, p.T278A, p.L281M, p.L281S, p.G285C, p.G285S, p.L295P, p.G296S, p.G321S, and p.R433W [52,54,58,68,73,74,76]. However, these studies would not rule out the possibility that some missense variants are also cytotoxic. Not a Kv7.4 variant, but the expression of a missense Kv7.2 variant, Kv7.2M518V, was reported to induce neuronal death [123]. Missense mutations that were reported to impair cell membrane targeting of Kv7.4, i.e., p.L274H, p.W276S, p.L281S, p.G285C, p.G285S, p.G296S, and p.G321S [58,68,81], are of particular interest because impaired cell membrane targeting implies a structural defect in a mutated protein, an intracellular accumulation of which could induce cellular stress. It should be pointed out that many previous studies reporting a dominant-negative inhibitory effect for Kv7 variants used transiently transfected cells. It is possible that a dominant-negative effect may be erroneously judged for some cytotoxic Kv7 variants that may indirectly affect the overall Kv7 channel activity by impairing cell viability, rather than directly affecting the function of Kv7WT. In fact, our whole-cell patch clamp recordings found “apparent” dominant-negative inhibitory effects in cells expressing both Kv7.4WT and Kv7.4242X or Kv7.4A349Pfs, which was most likely attributable to impaired cell membrane integrity [41]. Hence, cytotoxicity should be suspected and explored in any variants that affect the amino acid sequence of the Kv7 proteins (Table 1 and Table 2). The use of inducible stable cell lines is strongly encouraged to confidently identify cytotoxic variants and to quantify their relative cytotoxicity.
Since loss of function likely accounts for the pathological role of Kv7.1 variants underlying recessively inherited JLNS, it may seem insignificant to explore cytotoxicity in JLNS-associated Kv7.1 variants. In fact, our recent study identified WT-like small cytotoxicity in a JLNS-associated Kv7.1 variant, Kv7.1Q530X [41]. However, the generality of this finding should be thoroughly examined, especially given the observation that a recessively inherited Kv7.4 variant, Kv7.4A349Pfs, exhibits counterintuitively large cell-death-inducing cytotoxicity [41]. The relatively large number of JLNS-associated Kv7.1 variants (Table 1) allows for a systematic study examining the anticipated absence of large cytotoxicity that may account for their recessive inheritance.
Due to the remarkable structural similarity between Kv7.1 and Kv7.4 (Figure 1), a missense change of a conserved residue that affects the function of one Kv7 protein is expected to similarly affect the other. The fact that an approximately ten times greater number of disease-associated variants are found in KCNQ1 compared to KCNQ4 implies that the cardiac function is even sensitive to very small reductions in Kv7.1 channel activity. If true, why do JLNS-associated Kv7.1 variants, which induce severe cardiac (and hearing) phenotypes when they are homozygous or compound heterozygous, not induce haploinsufficiency-related cardiac phenotypes when heterozygous? In addition to examining cytotoxicity, it would be also worthwhile to learn if JLNS-associated Kv7.1 variants, especially missense ones, are functional by themselves and capable of physically interacting with Kv7.1WT to exert an inhibitory or cooperative effect. These experimental efforts may contribute to the determination of the magnitude of Kv7.1-mediated K+ channel activity that is minimally required to support normal cardiac and auditory functions.
It should be noted that “disease-associated” variants are not necessarily pathogenic. Experimental characterization is important for any disease-associated variants to define their pathogenicity. It is entirely possible that some variants have been falsely identified as pathogenic or likely pathogenic. Identification of nonpathogenic variants is as important as identification of truly pathogenic variants. The significance of experimental validation should not be understated regardless of the likelihood of a variant being pathogenic or nonpathogenic.

9. Concluding Remarks

Cytotoxicity-based pathological mechanisms have been acknowledged to account for several neurodegenerative diseases, such as Alzheimer’s disease, in which pathological aggregations of the tau protein are recognized as one of the hallmarks of the diseases [124], with over 100 dominantly inherited tau variants reported to date. However, cytotoxicity-based mechanisms have barely, if not never, been entertained for ion-channel-related hereditary hearing loss. Our efforts to experimentally confirm the absence of K+ channel activity in truncated Kv7.4 variants lacking the C-terminal tetramerization region might be seen as belaboring, given the fully established fact that tetramer formation is essential for completing a functional channel pore in any Kv channel. In fact, all three truncated Kv7.4 variants examined in our recent study were found to be nonfunctional by themselves, just as expected. However, cell-death-inducing cytotoxicity would have not been found if we did not pursue this thoroughly using inducible stable cell lines, underscoring the importance of experimentally characterizing disease-associated variants without preconceptions.
We identified large cell-death-inducing cytotoxicity in truncated Kv7.1 and Kv7.4 variants. However, as stressed above, cytotoxicity should be suspected in any variants, especially for dominantly inherited ones, that affect amino acid sequences of the Kv7 proteins. The currently pursued pharmacological strategy to augment residual Kv7 channel activity would not benefit patients with cytotoxic Kv7 variants. Thus, cytotoxic variants must be identified and distinguished from the others so that legitimate clinical strategies can be developed separately.

Funding

Deafness-related research in the author’s laboratory is supported by an NIH grant, DC017482.

Institutional Review Board Statement

All mouse procedures and protocols were approved by Northwestern University’s Institutional Animal Care and Use Committee (protocol: IS00022328).

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The author is extremely grateful to the Hugh Knowles Center for financially supporting an ongoing effort to generate several Kcnq4 knock-in mouse models.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Maljevic, S.; Wuttke, T.V.; Seebohm, G.; Lerche, H. KV7 channelopathies. Pflug. Arch. 2010, 460, 277–288. [Google Scholar] [CrossRef] [PubMed]
  2. Wang, J.J.; Li, Y. KCNQ potassium channels in sensory system and neural circuits. Acta Pharmacol. Sin. 2016, 37, 25–33. [Google Scholar] [CrossRef] [PubMed]
  3. Wang, Q.; Curran, M.E.; Splawski, I.; Burn, T.C.; Millholland, J.M.; VanRaay, T.J.; Shen, J.; Timothy, K.W.; Vincent, G.M.; de Jager, T.; et al. Positional cloning of a novel potassium channel gene: KVLQT1 mutations cause cardiac arrhythmias. Nat. Genet. 1996, 12, 17–23. [Google Scholar] [CrossRef] [PubMed]
  4. Neyroud, N.; Tesson, F.; Denjoy, I.; Leibovici, M.; Donger, C.; Barhanin, J.; Faure, S.; Gary, F.; Coumel, P.; Petit, C.; et al. A novel mutation in the potassium channel gene KVLQT1 causes the Jervell and Lange-Nielsen cardioauditory syndrome. Nat. Genet. 1997, 15, 186–189. [Google Scholar] [CrossRef] [PubMed]
  5. Westenskow, P.; Splawski, I.; Timothy, K.W.; Keating, M.T.; Sanguinetti, M.C. Compound mutations: A common cause of severe long-QT syndrome. Circulation 2004, 109, 1834–1841. [Google Scholar] [CrossRef]
  6. Priori, S.G.; Schwartz, P.J.; Napolitano, C.; Bianchi, L.; Dennis, A.; De Fusco, M.; Brown, A.M.; Casari, G. A recessive variant of the Romano-Ward long-QT syndrome? Circulation 1998, 97, 2420–2425. [Google Scholar] [CrossRef]
  7. Lang, F.; Vallon, V.; Knipper, M.; Wangemann, P. Functional significance of channels and transporters expressed in the inner ear and kidney. Am. J. Physiol. Cell Physiol. 2007, 293, C1187–C1208. [Google Scholar] [CrossRef]
  8. Beisel, K.W.; Nelson, N.C.; Delimont, D.C.; Fritzsch, B. Longitudinal gradients of KCNQ4 expression in spiral ganglion and cochlear hair cells correlate with progressive hearing loss in DFNA2. Mol. Brain Res. 2000, 82, 137–149. [Google Scholar] [CrossRef]
  9. Kharkovets, T.; Hardelin, J.P.; Safieddine, S.; Schweizer, M.; El-Amraoui, A.; Petit, C.; Jentsch, T.J. KCNQ4, a K+ channel mutated in a form of dominant deafness, is expressed in the inner ear and the central auditory pathway. Proc. Natl. Acad. Sci. USA 2000, 97, 4333–4338. [Google Scholar] [CrossRef]
  10. Brownell, W.E.; Bader, C.R.; Bertrand, D.; de Ribaupierre, Y. Evoked mechanical responses of isolated cochlear outer hair cells. Science 1985, 227, 194–196. [Google Scholar] [CrossRef]
  11. Evans, B.N.; Dallos, P. Stereocilia displacement induced somatic motility of cochlear outer hair cells. Proc. Natl. Acad. Sci. USA 1993, 90, 8347–8351. [Google Scholar] [CrossRef] [PubMed]
  12. Johnson, S.L.; Beurg, M.; Marcotti, W.; Fettiplace, R. Prestin-driven cochlear amplification is not limited by the outer hair cell membrane time constant. Neuron 2011, 70, 1143–1154. [Google Scholar] [CrossRef] [PubMed]
  13. Dominguez, L.M.; Dodson, K.M. Genetics of hearing loss: Focus on DFNA2. Appl. Clin. Genet. 2012, 5, 97–104. [Google Scholar] [CrossRef] [PubMed]
  14. Stenson, P.D.; Mort, M.; Ball, E.V.; Evans, K.; Hayden, M.; Heywood, S.; Hussain, M.; Phillips, A.D.; Cooper, D.N. The Human Gene Mutation Database: Towards a comprehensive repository of inherited mutation data for medical research, genetic diagnosis and next-generation sequencing studies. Hum. Genet. 2017, 136, 665–677. [Google Scholar] [CrossRef] [PubMed]
  15. Wangemann, P. Supporting sensory transduction: Cochlear fluid homeostasis and the endocochlear potential. J. Physiol. 2006, 576, 11–21. [Google Scholar] [CrossRef]
  16. Li, Y.; Liu, H.; Zhao, X.; He, D.Z. Endolymphatic Pote.ential Measured From Developing and Adult Mouse Inner Ear. Front. Cell Neurosci 2020, 14, 584928. [Google Scholar] [CrossRef]
  17. Lee, M.P.; Ravenel, J.D.; Hu, R.J.; Lustig, L.R.; Tomaselli, G.; Berger, R.D.; Brandenburg, S.A.; Litzi, T.J.; Bunton, T.E.; Limb, C.; et al. Targeted disruption of the Kvlqt1 gene causes deafness and gastric hyperplasia in mice. J. Clin. Investig. 2000, 106, 1447–1455. [Google Scholar] [CrossRef]
  18. Casimiro, M.C.; Knollmann, B.C.; Ebert, S.N.; Vary, J.C., Jr.; Greene, A.E.; Franz, M.R.; Grinberg, A.; Huang, S.P.; Pfeifer, K. Targeted disruption of the Kcnq1 gene produces a mouse model of Jervell and Lange-Nielsen Syndrome. Proc. Natl. Acad. Sci. USA 2001, 98, 2526–2531. [Google Scholar] [CrossRef]
  19. Rivas, A.; Francis, H.W. Inner ear abnormalities in a Kcnq1 (Kvlqt1) knockout mouse: A model of Jervell and Lange-Nielsen syndrome. Otol. Neurotol. 2005, 26, 415–424. [Google Scholar] [CrossRef]
  20. Chang, Q.; Wang, J.; Li, Q.; Kim, Y.; Zhou, B.; Wang, Y.; Li, H.; Lin, X. Virally mediated Kcnq1 gene replacement therapy in the immature scala media restores hearing in a mouse model of human Jervell and Lange-Nielsen deafness syndrome. EMBO Mol. Med. 2015, 7, 1077–1086. [Google Scholar] [CrossRef]
  21. Kharkovets, T.; Dedek, K.; Maier, H.; Schweizer, M.; Khimich, D.; Nouvian, R.; Vardanyan, V.; Leuwer, R.; Moser, T.; Jentsch, T.J. Mice with altered KCNQ4 K+ channels implicate sensory outer hair cells in human progressive deafness. EMBO J. 2006, 25, 642–652. [Google Scholar] [CrossRef] [PubMed]
  22. Carignano, C.; Barila, E.P.; Rias, E.I.; Dionisio, L.; Aztiria, E.; Spitzmaul, G. Inner Hair Cell and Neuron Degeneration Contribute to Hearing Loss in a DFNA2-Like Mouse Model. Neuroscience 2019, 410, 202–216. [Google Scholar] [CrossRef] [PubMed]
  23. Howard, R.J.; Clark, K.A.; Holton, J.M.; Minor, D.L., Jr. Structural insight into KCNQ (Kv7) channel assembly and channelopathy. Neuron 2007, 53, 663–675. [Google Scholar] [CrossRef] [PubMed]
  24. Wiener, R.; Haitin, Y.; Shamgar, L.; Fernandez-Alonso, M.C.; Martos, A.; Chomsky-Hecht, O.; Rivas, G.; Attali, B.; Hirsch, J.A. The KCNQ1 (Kv7.1) COOH terminus, a multitiered scaffold for subunit assembly and protein interaction. J. Biol. Chem. 2008, 283, 5815–5830. [Google Scholar] [CrossRef]
  25. Xu, Q.; Minor, D.L., Jr. Crystal structure of a trimeric form of the K(V)7.1 (KCNQ1) A-domain tail coiled-coil reveals structural plasticity and context dependent changes in a putative coiled-coil trimerization motif. Protein Sci. 2009, 18, 2100–2114. [Google Scholar] [CrossRef]
  26. Xu, Q.; Chang, A.; Tolia, A.; Minor, D.L., Jr. Structure of a Ca2+/CaM:Kv7.4 (KCNQ4) B-helix complex provides insight into M current modulation. J. Mol. Biol. 2013, 425, 378–394. [Google Scholar] [CrossRef]
  27. Sachyani, D.; Dvir, M.; Strulovich, R.; Tria, G.; Tobelaim, W.; Peretz, A.; Pongs, O.; Svergun, D.; Attali, B.; Hirsch, J.A. 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. [Google Scholar] [CrossRef]
  28. Sun, J.; MacKinnon, R. Cryo-EM Structure of a KCNQ1/CaM Complex Reveals Insights into Congenital Long QT Syndrome. Cell 2017, 169, 1042–1050 e1049. [Google Scholar] [CrossRef]
  29. Sun, J.; MacKinnon, R. Structural Basis of Human KCNQ1 Modulation and Gating. Cell 2020, 180, 340–347.e9. [Google Scholar] [CrossRef]
  30. Li, T.; Wu, K.; Yue, Z.; Wang, Y.; Zhang, F.; Shen, H. Structural Basis for the Modulation of Human KCNQ4 by Small-Molecule Drugs. Mol. Cell 2021, 81, 25–37.e4. [Google Scholar] [CrossRef]
  31. Zheng, Y.; Liu, H.; Chen, Y.; Dong, S.; Wang, F.; Wang, S.; Li, G.L.; Shu, Y.; Xu, F. Structural insights into the lipid and ligand regulation of a human neuronal KCNQ channel. Neuron 2022, 110, 237–247.e4. [Google Scholar] [CrossRef]
  32. Salata, J.J.; Jurkiewicz, N.K.; Wang, J.; Evans, B.E.; Orme, H.T.; Sanguinetti, M.C. A novel benzodiazepine that activates cardiac slow delayed rectifier K+ currents. Mol. Pharmacol. 1998, 54, 220–230. [Google Scholar] [CrossRef] [PubMed]
  33. Xu, X.; Salata, J.J.; Wang, J.; Wu, Y.; Yan, G.X.; Liu, T.; Marinchak, R.A.; Kowey, P.R. Increasing I(Ks) corrects abnormal repolarization in rabbit models of acquired LQT2 and ventricular hypertrophy. Am. J. Physiol. Heart Circ. Physiol. 2002, 283, H664–H670. [Google Scholar] [CrossRef] [PubMed]
  34. Wuttke, T.V.; Lerche, H. Novel anticonvulsant drugs targeting voltage-dependent ion channels. Expert Opin. Investig. Drugs 2006, 15, 1167–1177. [Google Scholar] [CrossRef] [PubMed]
  35. Xiong, Q.; Sun, H.; Li, M. Zinc pyrithione-mediated activation of voltage-gated KCNQ potassium channels rescues epileptogenic mutants. Nat. Chem. Biol. 2007, 3, 287–296. [Google Scholar] [CrossRef]
  36. Leitner, M.G.; Feuer, A.; Ebers, O.; Schreiber, D.N.; Halaszovich, C.R.; Oliver, D. Restoration of ion channel function in deafness-causing KCNQ4 mutants by synthetic channel openers. Br. J. Pharmacol. 2012, 165, 2244–2259. [Google Scholar] [CrossRef]
  37. Tian, C.; Zhu, R.; Zhu, L.; Qiu, T.; Cao, Z.; Kang, T. Potassium channels: Structures, diseases, and modulators. Chem. Biol. Drug Des. 2014, 83, 1–26. [Google Scholar] [CrossRef]
  38. Imbrici, P.; Liantonio, A.; Camerino, G.M.; De Bellis, M.; Camerino, C.; Mele, A.; Giustino, A.; Pierno, S.; De Luca, A.; Tricarico, D.; et al. Therapeutic Approaches to Genetic Ion Channelopathies and Perspectives in Drug Discovery. Front. Pharmacol. 2016, 7, 121. [Google Scholar] [CrossRef]
  39. Miceli, F.; Soldovieri, M.V.; Ambrosino, P.; Manocchio, L.; Mosca, I.; Taglialatela, M. Pharmacological Targeting of Neuronal Kv7.2/3 Channels: A Focus on Chemotypes and Receptor Sites. Curr. Med. Chem. 2018, 25, 2637–2660. [Google Scholar] [CrossRef]
  40. Barrese, V.; Stott, J.B.; Greenwood, I.A. KCNQ-Encoded Potassium Channels as Therapeutic Targets. Annu. Rev. Pharmacol. Toxicol. 2018, 58, 625–648. [Google Scholar] [CrossRef]
  41. Shin, D.H.; Jung, J.; Koh, Y.I.; Rim, J.H.; Lee, J.S.; Choi, H.J.; Joo, S.Y.; Yu, S.; Cha, D.H.; Lee, S.Y.; et al. A recurrent mutation in KCNQ4 in Korean families with nonsyndromic hearing loss and rescue of the channel activity by KCNQ activators. Hum. Mutat. 2019, 40, 335–346. [Google Scholar] [CrossRef] [PubMed]
  42. Yu, H.; Wu, M.; Townsend, S.D.; Zou, B.; Long, S.; Daniels, J.S.; McManus, O.B.; Li, M.; Lindsley, C.W.; Hopkins, C.R. Discovery, Synthesis, and Structure Activity Relationship of a Series of N-Aryl- bicyclo[2.2.1]heptane-2-carboxamides: Characterization of ML213 as a Novel KCNQ2 and KCNQ4 Potassium Channel Opener. ACS Chem. Neurosci. 2011, 2, 572–577. [Google Scholar] [CrossRef]
  43. Wang, J.; Liu, Y.; Hu, F.; Yang, J.; Guo, X.; Hou, X.; Ju, C.; Wang, K. Activation of Neuronal Voltage-Gated Potassium Kv7/KCNQ/M-Current by a Novel Channel Opener SCR2682 for Alleviation of Chronic Pain. J. Pharmacol. Exp. Ther. 2021, 377, 20–28. [Google Scholar] [CrossRef] [PubMed]
  44. Kamada, F.; Kure, S.; Kudo, T.; Suzuki, Y.; Oshima, T.; Ichinohe, A.; Kojima, K.; Niihori, T.; Kanno, J.; Narumi, Y.; et al. A novel KCNQ4 one-base deletion in a large pedigree with hearing loss: Implication for the genotype-phenotype correlation. J. Hum. Genet. 2006, 51, 455–460. [Google Scholar] [CrossRef] [PubMed]
  45. Hildebrand, M.S.; Tack, D.; McMordie, S.J.; DeLuca, A.; Hur, I.A.; Nishimura, C.; Huygen, P.; Casavant, T.L.; Smith, R.J. Audioprofile-directed screening identifies novel mutations in KCNQ4 causing hearing loss at the DFNA2 locus. Genet. Med. 2008, 10, 797–804. [Google Scholar] [CrossRef]
  46. Wasano, K.; Mutai, H.; Obuchi, C.; Masuda, S.; Matsunaga, T. A novel frameshift mutation in KCNQ4 in a family with autosomal recessive non-syndromic hearing loss. Biochem. Biophys. Res. Commun. 2015, 463, 582–586. [Google Scholar] [CrossRef]
  47. Kojima, T.; Wasano, K.; Takahashi, S.; Homma, K. Cell death-inducing cytotoxicity in truncated KCNQ4 variants associated with DFNA2 hearing loss. Dis. Model. Mech. 2021, 14, dmm049015. [Google Scholar] [CrossRef]
  48. Tranebjaerg, L.; Bathen, J.; Tyson, J.; Bitner-Glindzicz, M. Jervell and Lange-Nielsen syndrome: A Norwegian perspective. Am. J. Med. Genet. 1999, 89, 137–146. [Google Scholar] [CrossRef]
  49. Huang, L.; Bitner-Glindzicz, M.; Tranebjaerg, L.; Tinker, A. A spectrum of functional effects for disease causing mutations in the Jervell and Lange-Nielsen syndrome. Cardiovasc. Res. 2001, 51, 670–680. [Google Scholar] [CrossRef]
  50. Wilson, A.J.; Quinn, K.V.; Graves, F.M.; Bitner-Glindzicz, M.; Tinker, A. Abnormal KCNQ1 trafficking influences disease pathogenesis in hereditary long QT syndromes (LQT1). Cardiovasc. Res. 2005, 67, 476–486. [Google Scholar] [CrossRef] [Green Version]
  51. Davis, S.N.; Wu, P.; Camci, E.D.; Simon, J.A.; Rubel, E.W.; Raible, D.W. Chloroquine kills hair cells in zebrafish lateral line and murine cochlear cultures: Implications for ototoxicity. Hear. Res. 2020, 395, 108019. [Google Scholar] [CrossRef] [PubMed]
  52. Vincent, G.M.; Timothy, K.W.; Leppert, M.; Keating, M. The spectrum of symptoms and QT intervals in carriers of the gene for the long-QT syndrome. N. Engl. J. Med. 1992, 327, 846–852. [Google Scholar] [CrossRef] [PubMed]
  53. Vincent, G.M. Long QT syndrome. Cardiol. Clin. 2000, 18, 309–325. [Google Scholar] [CrossRef]
  54. Priori, S.G.; Napolitano, C.; Schwartz, P.J. Low penetrance in the long-QT syndrome: Clinical impact. Circulation 1999, 99, 529–533. [Google Scholar] [CrossRef] [PubMed]
  55. Chen, S.; Zhang, L.; Bryant, R.M.; Vincent, G.M.; Flippin, M.; Lee, J.C.; Brown, E.; Zimmerman, F.; Rozich, R.; Szafranski, P.; et al. KCNQ1 mutations in patients with a family history of lethal cardiac arrhythmias and sudden death. Clin. Genet. 2003, 63, 273–282. [Google Scholar] [CrossRef]
  56. Attili, V.S.; Bapsy, P.P.; Anupama, G.; Lokanatha, D. Irreversible sensorineural hearing loss due to Imatinib. Leuk. Res. 2008, 32, 991–992. [Google Scholar] [CrossRef]
  57. Wasif, K.; Wasif, N.; Saif, M.W. Imatinib-induced Ototoxicity in a Patient with Gastrointestinal Stromal Tumor (GIST). Cureus 2016, 8, e848. [Google Scholar] [CrossRef]
  58. Kubisch, C.; Schroeder, B.C.; Friedrich, T.; Lutjohann, B.; El-Amraoui, A.; Marlin, S.; Petit, C.; Jentsch, T.J. KCNQ4, a novel potassium channel expressed in sensory outer hair cells, is mutated in dominant deafness. Cell 1999, 96, 437–446. [Google Scholar] [CrossRef]
  59. Baek, J.I.; Park, H.J.; Park, K.; Choi, S.J.; Lee, K.Y.; Yi, J.H.; Friedman, T.B.; Drayna, D.; Shin, K.S.; Kim, U.K. Pathogenic effects of a novel mutation (c.664_681del) in KCNQ4 channels associated with auditory pathology. Biochim. Biophys. Acta 2011, 1812, 536–543. [Google Scholar] [CrossRef]
  60. Kim, H.J.; Lv, P.; Sihn, C.R.; Yamoah, E.N. Cellular and molecular mechanisms of autosomal dominant form of progressive hearing loss, DFNA2. J. Biol. Chem. 2011, 286, 1517–1527. [Google Scholar] [CrossRef] [Green Version]
  61. Gao, Y.; Yechikov, S.; Vazquez, A.E.; Chen, D.; Nie, L. Impaired surface expression and conductance of the KCNQ4 channel lead to sensorineural hearing loss. J. Cell. Mol. Med. 2013, 17, 889–900. [Google Scholar] [CrossRef] [PubMed]
  62. Jung, J.; Lin, H.; Koh, Y.I.; Ryu, K.; Lee, J.S.; Rim, J.H.; Choi, H.J.; Lee, H.J.; Kim, H.Y.; Yu, S.; et al. Rare KCNQ4 variants found in public databases underlie impaired channel activity that may contribute to hearing impairment. Exp. Mol. Med. 2019, 51, 1–12. [Google Scholar] [CrossRef] [PubMed]
  63. Kim, E.C.; Zhang, J.; Pang, W.; Wang, S.; Lee, K.Y.; Cavaretta, J.P.; Walters, J.; Procko, E.; Tsai, N.P.; Chung, H.J. Reduced axonal surface expression and phosphoinositide sensitivity in Kv7 channels disrupts their function to inhibit neuronal excitability in Kcnq2 epileptic encephalopathy. Neurobiol. Dis. 2018, 118, 76–93. [Google Scholar] [CrossRef] [PubMed]
  64. Mencia, A.; Gonzalez-Nieto, D.; Modamio-Hoybjor, S.; Etxeberria, A.; Aranguez, G.; Salvador, N.; Del Castillo, I.; Villarroel, A.; Moreno, F.; Barrio, L.; et al. A novel KCNQ4 pore-region mutation (p.G296S) causes deafness by impairing cell-surface channel expression. Hum. Genet. 2008, 123, 41–53. [Google Scholar] [CrossRef] [PubMed]
  65. Zhang, Y.; Wu, K.M.; Yang, L.; Dong, Q.; Yu, J.T. Tauopathies: New perspectives and challenges. Mol. Neurodegener. 2022, 17, 28. [Google Scholar] [CrossRef]
  66. Wang, R.R.; Li, N.; Zhang, Y.H.; Wang, L.L.; Teng, S.Y.; Pu, J.L. Novel compound heterozygous mutations T2C and 1149insT in the KCNQ1 gene cause Jervell and Lange-Nielsen syndrome. Int. J. Mol. Med. 2011, 28, 41–46. [Google Scholar] [CrossRef]
  67. Nishimura, M.; Ueda, M.; Ebata, R.; Utsuno, E.; Ishii, T.; Matsushita, K.; Ohara, O.; Shimojo, N.; Kobayashi, Y.; Nomura, F. A novel KCNQ1 nonsense variant in the isoform-specific first exon causes both jervell and Lange-Nielsen syndrome 1 and long QT syndrome 1: A case report. BMC Med. Genet. 2017, 18, 66. [Google Scholar] [CrossRef]
  68. Wang, C.; Lu, Y.; Cheng, J.; Zhang, L.; Liu, W.; Peng, W.; Zhang, D.; Duan, H.; Han, D.; Yuan, H. Identification of KCNQ1 compound heterozygous mutations in three Chinese families with Jervell and Lange-Nielsen Syndrome. Acta Otolaryngol. 2017, 137, 522–528. [Google Scholar] [CrossRef]
  69. Vyas, B.; Puri, R.D.; Namboodiri, N.; Nair, M.; Sharma, D.; Movva, S.; Saxena, R.; Bohora, S.; Aggarwal, N.; Vora, A.; et al. KCNQ1 mutations associated with Jervell and Lange-Nielsen syndrome and autosomal recessive Romano-Ward syndrome in India-expanding the spectrum of long QT syndrome type 1. Am. J. Med. Genet. A 2016, 170, 1510–1519. [Google Scholar] [CrossRef]
  70. Wang, Z.; Li, H.; Moss, A.J.; Robinson, J.; Zareba, W.; Knilans, T.; Bowles, N.E.; Towbin, J.A. Compound heterozygous mutations in KvLQT1 cause Jervell and Lange-Nielsen syndrome. Mol. Genet. Metab. 2002, 75, 308–316. [Google Scholar] [CrossRef]
  71. Eldstrom, J.; Xu, H.; Werry, D.; Kang, C.; Loewen, M.E.; Degenhardt, A.; Sanatani, S.; Tibbits, G.F.; Sanders, C.; Fedida, D. Mechanistic basis for LQT1 caused by S3 mutations in the KCNQ1 subunit of IKs. J. Gen. Physiol. 2010, 135, 433–448. [Google Scholar] [CrossRef]
  72. Mohammad-Panah, R.; Demolombe, S.; Neyroud, N.; Guicheney, P.; Kyndt, F.; van den Hoff, M.; Baro, I.; Escande, D. Mutations in a dominant-negative isoform correlate with phenotype in inherited cardiac arrhythmias. Am. J. Hum. Genet. 1999, 64, 1015–1023. [Google Scholar] [CrossRef] [PubMed]
  73. Chouabe, C.; Neyroud, N.; Richard, P.; Denjoy, I.; Hainque, B.; Romey, G.; Drici, M.D.; Guicheney, P.; Barhanin, J. Novel mutations in KvLQT1 that affect Iks activation through interactions with Isk. Cardiovasc. Res. 2000, 45, 971–980. [Google Scholar] [CrossRef]
  74. Park, K.H.; Piron, J.; Dahimene, S.; Merot, J.; Baro, I.; Escande, D.; Loussouarn, G. Impaired KCNQ1-KCNE1 and phosphatidylinositol-4,5-bisphosphate interaction underlies the long QT syndrome. Circ. Res. 2005, 96, 730–739. [Google Scholar] [CrossRef] [PubMed]
  75. Itoh, H.; Sakaguchi, T.; Ding, W.G.; Watanabe, E.; Watanabe, I.; Nishio, Y.; Makiyama, T.; Ohno, S.; Akao, M.; Higashi, Y.; et al. Latent genetic backgrounds and molecular pathogenesis in drug-induced long-QT syndrome. Circ. Arrhythm. Electrophysiol. 2009, 2, 511–523. [Google Scholar] [CrossRef]
  76. Coyan, F.C.; Abderemane-Ali, F.; Amarouch, M.Y.; Piron, J.; Mordel, J.; Nicolas, C.S.; Steenman, M.; Merot, J.; Marionneau, C.; Thomas, A.; et al. A long QT mutation substitutes cholesterol for phosphatidylinositol-4,5-bisphosphate in KCNQ1 channel regulation. PLoS ONE 2014, 9, e93255. [Google Scholar] [CrossRef]
  77. Coto, E.; Garcia-Fernandez, F.J.; Calvo, D.; Salgado-Aranda, R.; Martin-Gonzalez, J.; Alonso, B.; Iglesias, S.; Gomez, J. An elderly Jervell and Lange-Nielsen patient heterozygous compound for two new KCNQ1 mutations. Am. J. Med. Genet. A 2017, 173, 749–752. [Google Scholar] [CrossRef]
  78. Tyson, J.; Tranebjaerg, L.; McEntagart, M.; Larsen, L.A.; Christiansen, M.; Whiteford, M.L.; Bathen, J.; Aslaksen, B.; Sorland, S.J.; Lund, O.; et al. Mutational spectrum in the cardioauditory syndrome of Jervell and Lange-Nielsen. Hum. Genet. 2000, 107, 499–503. [Google Scholar] [CrossRef]
  79. Neyroud, N.; Denjoy, I.; Donger, C.; Gary, F.; Villain, E.; Leenhardt, A.; Benali, K.; Schwartz, K.; Coumel, P.; Guicheney, P. Heterozygous mutation in the pore of potassium channel gene KvLQT1 causes an apparently normal phenotype in long QT syndrome. Eur. J. Hum. Genet. 1998, 6, 129–133. [Google Scholar] [CrossRef]
  80. Chouabe, C.; Neyroud, N.; Guicheney, P.; Lazdunski, M.; Romey, G.; Barhanin, J. Properties of KvLQT1 K+ channel mutations in Romano-Ward and Jervell and Lange-Nielsen inherited cardiac arrhythmias. EMBO J. 1997, 16, 5472–5479. [Google Scholar] [CrossRef] [Green Version]
  81. Matsuda, S.; Ohnuki, Y.; Okami, M.; Ochiai, E.; Yamada, S.; Takahashi, K.; Osawa, M.; Okami, K.; Iida, M.; Mochizuki, H. Jervell and Lange-Nielsen syndrome with novel KCNQ1 and additional gene mutations. Hum. Genome Var. 2020, 7, 34. [Google Scholar] [CrossRef] [PubMed]
  82. Zhang, Y.; Li, X.; Yang, Y.; Wang, J.; Gao, X.; Fan, M. The combined novel KCNQ1 frameshift I145Sfs*92 and nonsense W392X variants caused Jervell and Lange-Nielsen syndrome in a Chinese infant presenting with sustained foetal bradycardia. Europace 2020, 22, 1880–1884. [Google Scholar] [CrossRef] [PubMed]
  83. Harmer, S.C.; Mohal, J.S.; Royal, A.A.; McKenna, W.J.; Lambiase, P.D.; Tinker, A. Cellular mechanisms underlying the increased disease severity seen for patients with long QT syndrome caused by compound mutations in KCNQ1. Biochem. J. 2014, 462, 133–142. [Google Scholar] [CrossRef] [PubMed]
  84. Qiu, Y.; Chen, S.; Wu, X.; Zhang, W.J.; Xie, W.; Jin, Y.; Xie, L.; Xu, K.; Bai, X.; Zhang, H.M.; et al. Jervell and Lange-Nielsen Syndrome due to a Novel Compound Heterozygous KCNQ1 Mutation in a Chinese Family. Neural. Plast. 2020, 2020, 3569359. [Google Scholar] [CrossRef]
  85. Chen, Q.; Zhang, D.; Gingell, R.L.; Moss, A.J.; Napolitano, C.; Priori, S.G.; Schwartz, P.J.; Kehoe, E.; Robinson, J.L.; Schulze-Bahr, E.; et al. Homozygous deletion in KVLQT1 associated with Jervell and Lange-Nielsen syndrome. Circulation 1999, 99, 1344–1347. [Google Scholar] [CrossRef]
  86. Amirian, A.; Zafari, Z.; Dalili, M.; Saber, S.; Karimipoor, M.; Dabbagh Bagheri, S.; Fazelifar, A.F.; Zeinali, S. Detection of a new KCNQ1 frameshift mutation associated with Jervell and Lange-Nielsen syndrome in 2 Iranian families. J. Arrhythm. 2018, 34, 286–290. [Google Scholar] [CrossRef]
  87. Al-Aama, J.Y.; Al-Ghamdi, S.; Bdier, A.Y.; Wilde, A.A.; Bhuiyan, Z.A. De novo mutation in the KCNQ1 gene causal to Jervell and Lange-Nielsen syndrome. Clin. Genet. 2014, 86, 492–495. [Google Scholar] [CrossRef]
  88. Chung, S.K.; MacCormick, J.M.; McCulley, C.H.; Crawford, J.; Eddy, C.A.; Mitchell, E.A.; Shelling, A.N.; French, J.K.; Skinner, J.R.; Rees, M.I. Long QT and Brugada syndrome gene mutations in New Zealand. Heart Rhythm. 2007, 4, 1306–1314. [Google Scholar] [CrossRef]
  89. Wei, J.; Fish, F.A.; Myerburg, R.J.; Roden, D.M.; George, A.L., Jr. Novel KCNQ1 mutations associated with recessive and dominant congenital long QT syndromes: Evidence for variable hearing phenotype associated with R518X. Hum. Mutat. 2000, 15, 387–388. [Google Scholar] [CrossRef]
  90. Gao, Y.; Li, C.; Liu, W.; Wu, R.; Qiu, X.; Liang, R.; Li, L.; Zhang, L.; Hu, D. Genotype-phenotype analysis of three Chinese families with Jervell and Lange-Nielsen syndrome. J. Cardiovasc. Dis. Res. 2012, 3, 67–75. [Google Scholar] [CrossRef] [Green Version]
  91. Vojdani, S.; Amirsalari, S.; Milanizadeh, S.; Molaei, F.; Ajalloueyane, M.; Khosravi, A.; Hamzehzadeh, L.; Ghasemi, M.M.; Talee, M.R.; Abbaszadegan, M.R. Mutation Screening of KCNQ1 and KCNE1 Genes in Iranian Patients With Jervell and Lange-Nielsen Syndrome. Fetal Pediatr. Pathol. 2019, 38, 273–281. [Google Scholar] [CrossRef] [PubMed]
  92. Splawski, I.; Timothy, K.W.; Vincent, G.M.; Atkinson, D.L.; Keating, M.T. Molecular basis of the long-QT syndrome associated with deafness. N. Engl. J. Med. 1997, 336, 1562–1567. [Google Scholar] [CrossRef] [PubMed]
  93. Ohno, S.; Kubota, T.; Yoshida, H.; Tsuji, K.; Makiyama, T.; Yamada, S.; Kuga, K.; Yamaguchi, I.; Kita, T.; Horie, M. A novel mutation associated with Jervell and Lange-Nielsen syndrome in a Japanese family. Circ. J. 2008, 72, 687–693. [Google Scholar] [CrossRef] [PubMed]
  94. Safka Brozkova, D.; Uhrova Meszarosova, A.; Lassuthova, P.; Varga, L.; Stanek, D.; Borecka, S.; Lastuvkova, J.; Cejnova, V.; Raskova, D.; Lhota, F.; et al. The Cause of Hereditary Hearing Loss in GJB2 Heterozygotes-A Comprehensive Study of the GJB2/DFNB1 Region. Genes 2021, 12, 684. [Google Scholar] [CrossRef]
  95. Iwasa, Y.I.; Nishio, S.Y.; Usami, S.I. Comprehensive Genetic Analysis of Japanese Autosomal Dominant Sensorineural Hearing Loss Patients. PLoS ONE 2016, 11, e0166781. [Google Scholar] [CrossRef]
  96. Su, C.C.; Yang, J.J.; Shieh, J.C.; Su, M.C.; Li, S.Y. Identification of novel mutations in the KCNQ4 gene of patients with nonsyndromic deafness from Taiwan. Audiol. Neurootol. 2007, 12, 20–26. [Google Scholar] [CrossRef]
  97. Walls, W.D.; Moteki, H.; Thomas, T.R.; Nishio, S.Y.; Yoshimura, H.; Iwasa, Y.; Frees, K.L.; Nishimura, C.J.; Azaiez, H.; Booth, K.T.; et al. A comparative analysis of genetic hearing loss phenotypes in European/American and Japanese populations. Hum. Genet. 2020, 139, 1315–1323. [Google Scholar] [CrossRef]
  98. Sloan-Heggen, C.M.; Bierer, A.O.; Shearer, A.E.; Kolbe, D.L.; Nishimura, C.J.; Frees, K.L.; Ephraim, S.S.; Shibata, S.B.; Booth, K.T.; Campbell, C.A.; et al. Comprehensive genetic testing in the clinical evaluation of 1119 patients with hearing loss. Hum. Genet. 2016, 135, 441–450. [Google Scholar] [CrossRef]
  99. Naito, T.; Nishio, S.Y.; Iwasa, Y.; Yano, T.; Kumakawa, K.; Abe, S.; Ishikawa, K.; Kojima, H.; Namba, A.; Oshikawa, C.; et al. Comprehensive genetic screening of KCNQ4 in a large autosomal dominant nonsyndromic hearing loss cohort: Genotype-phenotype correlations and a founder mutation. PLoS ONE 2013, 8, e63231. [Google Scholar] [CrossRef]
  100. Xia, X.; Zhang, Q.; Jia, Y.; Shu, Y.; Yang, J.; Yang, H.; Yan, Z. Molecular basis and restoration of function deficiencies of Kv7.4 variants associated with inherited hearing loss. Hear. Res. 2020, 388, 107884. [Google Scholar] [CrossRef]
  101. Uehara, D.T.; Freitas, E.L.; Alves, L.U.; Mazzeu, J.F.; Auricchio, M.T.; Tabith, A., Jr.; Monteiro, M.L.; Rosenberg, C.; Mingroni-Netto, R.C. A novel KCNQ4 mutation and a private IMMP2L-DOCK4 duplication segregating with nonsyndromic hearing loss in a Brazilian family. Hum. Genome Var. 2015, 2, 15038. [Google Scholar] [CrossRef] [PubMed]
  102. Jung, J.; Choi, H.B.; Koh, Y.I.; Rim, J.H.; Choi, H.J.; Kim, S.H.; Lee, J.H.; An, J.; Kim, A.; Lee, J.S.; et al. Whole-exome sequencing identifies two novel mutations in KCNQ4 in individuals with nonsyndromic hearing loss. Sci. Rep. 2018, 8, 16659. [Google Scholar] [CrossRef] [PubMed]
  103. Namba, K.; Mutai, H.; Kaneko, H.; Hashimoto, S.; Matsunaga, T. In silico modeling of the pore region of a KCNQ4 missense mutant from a patient with hearing loss. BMC Res. Notes. 2012, 5, 145. [Google Scholar] [CrossRef] [PubMed]
  104. Van Hauwe, P.; Coucke, P.J.; Ensink, R.J.; Huygen, P.; Cremers, C.W.; Van Camp, G. Mutations in the KCNQ4 K+ channel gene, responsible for autosomal dominant hearing loss, cluster in the channel pore region. Am. J. Med. Genet. 2000, 93, 184–187. [Google Scholar] [CrossRef]
  105. Wang, H.; Zhao, Y.; Yi, Y.; Gao, Y.; Liu, Q.; Wang, D.; Li, Q.; Lan, L.; Li, N.; Guan, J.; et al. Targeted high-throughput sequencing identifies pathogenic mutations in KCNQ4 in two large Chinese families with autosomal dominant hearing loss. PLoS ONE 2014, 9, e103133. [Google Scholar] [CrossRef]
  106. Plevova, P.; Paprskarova, M.; Tvrda, P.; Turska, P.; Slavkovsky, R.; Mrazkova, E. STRC Deletion is a Frequent Cause of Slight to Moderate Congenital Hearing Impairment in the Czech Republic. Otol. Neurotol. 2017, 38, e393–e400. [Google Scholar] [CrossRef]
  107. Sommen, M.; Schrauwen, I.; Vandeweyer, G.; Boeckx, N.; Corneveaux, J.J.; van den Ende, J.; Boudewyns, A.; De Leenheer, E.; Janssens, S.; Claes, K.; et al. DNA Diagnostics of Hereditary Hearing Loss: A Targeted Resequencing Approach Combined with a Mutation Classification System. Hum. Mutat. 2016, 37, 812–819. [Google Scholar] [CrossRef]
  108. Coucke, P.J.; Van Hauwe, P.; Kelley, P.M.; Kunst, H.; Schatteman, I.; Van Velzen, D.; Meyers, J.; Ensink, R.J.; Verstreken, M.; Declau, F.; et al. Mutations in the KCNQ4 gene are responsible for autosomal dominant deafness in four DFNA2 families. Hum. Mol. Genet. 1999, 8, 1321–1328. [Google Scholar] [CrossRef]
  109. Talebizadeh, Z.; Kelley, P.M.; Askew, J.W.; Beisel, K.W.; Smith, S.D. Novel mutation in the KCNQ4 gene in a large kindred with dominant progressive hearing loss. Hum. Mutat. 1999, 14, 493–501. [Google Scholar] [CrossRef]
  110. Bal, M.; Zhang, J.; Zaika, O.; Hernandez, C.C.; Shapiro, M.S. Homomeric and heteromeric assembly of KCNQ (Kv7) K+ channels assayed by total internal reflection fluorescence/fluorescence resonance energy transfer and patch clamp analysis. J. Biol. Chem. 2008, 283, 30668–30676. [Google Scholar] [CrossRef] [Green Version]
  111. Smedley, D.; Smith, K.R.; Martin, A.; Thomas, E.A.; McDonagh, E.M.; Cipriani, V.; Ellingford, J.M.; Arno, G.; Tucci, A.; Vandrovcova, J.; et al. 100,000 Genomes Pilot on Rare-Disease Diagnosis in Health Care—Preliminary Report. N. Engl. J. Med. 2021, 385, 1868–1880. [Google Scholar] [CrossRef]
  112. Arnett, J.; Emery, S.B.; Kim, T.B.; Boerst, A.K.; Lee, K.; Leal, S.M.; Lesperance, M.M. Autosomal dominant progressive sensorineural hearing loss due to a novel mutation in the KCNQ4 gene. Arch. Otolaryngol. Head Neck Surg. 2011, 137, 54–59. [Google Scholar] [CrossRef] [PubMed]
  113. Huang, B.; Liu, Y.; Gao, X.; Xu, J.; Dai, P.; Zhu, Q.; Yuan, Y. A novel pore-region mutation, c.887G > A (p.G296D) in KCNQ4, causing hearing loss in a Chinese family with autosomal dominant non-syndromic deafness 2. BMC Med. Genet. 2017, 18, 36. [Google Scholar] [CrossRef] [PubMed]
  114. Morgan, A.; Lenarduzzi, S.; Cappellani, S.; Pecile, V.; Morgutti, M.; Orzan, E.; Ghiselli, S.; Ambrosetti, U.; Brumat, M.; Gajendrarao, P.; et al. Genomic Studies in a Large Cohort of Hearing Impaired Italian Patients Revealed Several New Alleles, a Rare Case of Uniparental Disomy (UPD) and the Importance to Search for Copy Number Variations. Front. Genet. 2018, 9, 681. [Google Scholar] [CrossRef] [PubMed]
  115. Lee, S.Y.; Choi, H.B.; Park, M.; Choi, I.S.; An, J.; Kim, A.; Kim, E.; Kim, N.; Han, J.H.; Kim, M.Y.; et al. Novel KCNQ4 variants in different functional domains confer genotype- and mechanism-based therapeutics in patients with nonsyndromic hearing loss. Exp. Mol. Med. 2021, 53, 1192–1204. [Google Scholar] [CrossRef] [PubMed]
  116. Boucher, S.; Tai, F.W.J.; Delmaghani, S.; Lelli, A.; Singh-Estivalet, A.; Dupont, T.; Niasme-Grare, M.; Michel, V.; Wolff, N.; Bahloul, A.; et al. Ultrarare heterozygous pathogenic variants of genes causing dominant forms of early-onset deafness underlie severe presbycusis. Proc. Natl. Acad. Sci. USA 2020, 117, 31278–31289. [Google Scholar] [CrossRef]
  117. Khan, A.; Han, S.; Wang, R.; Ansar, M.; Ahmad, W.; Zhang, X. Sequence variants in genes causing nonsyndromic hearing loss in a Pakistani cohort. Mol. Genet. Genom. Med. 2019, 7, e917. [Google Scholar] [CrossRef]
  118. Van Laer, L.; Carlsson, P.I.; Ottschytsch, N.; Bondeson, M.L.; Konings, A.; Vandevelde, A.; Dieltjens, N.; Fransen, E.; Snyders, D.; Borg, E.; et al. The contribution of genes involved in potassium-recycling in the inner ear to noise-induced hearing loss. Hum. Mutat. 2006, 27, 786–795. [Google Scholar] [CrossRef]
  119. Mehregan, H.; Mohseni, M.; Akbari, M.; Jalalvand, K.; Arzhangi, S.; Nikzat, N.; Kahrizi, K.; Najmabadi, H. Novel Mutations in KCNQ4, LHFPL5 and COCH Genes in Iranian Families with Hearing Impairment. Arch. Iran. Med. 2019, 22, 189–197. [Google Scholar]
  120. Miyagawa, M.; Naito, T.; Nishio, S.Y.; Kamatani, N.; Usami, S. Targeted exon sequencing successfully discovers rare causative genes and clarifies the molecular epidemiology of Japanese deafness patients. PLoS ONE 2013, 8, e71381. [Google Scholar] [CrossRef]
  121. Wu, C.C.; Lin, Y.H.; Lu, Y.C.; Chen, P.J.; Yang, W.S.; Hsu, C.J.; Chen, P.L. Application of massively parallel sequencing to genetic diagnosis in multiplex families with idiopathic sensorineural hearing impairment. PLoS ONE 2013, 8, e57369. [Google Scholar] [CrossRef] [PubMed]
  122. Watabe, T.; Matsunaga, T.; Namba, K.; Mutai, H.; Inoue, Y.; Ogawa, K. Moderate hearing loss associated with a novel KCNQ4 non-truncating mutation located near the N-terminus of the pore helix. Biochem. Biophys. Res. Commun. 2013, 432, 475–479. [Google Scholar] [CrossRef] [PubMed]
  123. Johnston, J.J.; Lewis, K.L.; Ng, D.; Singh, L.N.; Wynter, J.; Brewer, C.; Brooks, B.P.; Brownell, I.; Candotti, F.; Gonsalves, S.G.; et al. Individualized iterative phenotyping for genome-wide analysis of loss-of-function mutations. Am. J. Hum. Genet. 2015, 96, 913–925. [Google Scholar] [CrossRef]
  124. Ceyhan-Birsoy, O.; Murry, J.B.; Machini, K.; Lebo, M.S.; Yu, T.W.; Fayer, S.; Genetti, C.A.; Schwartz, T.S.; Agrawal, P.B.; Parad, R.B.; et al. Interpretation of Genomic Sequencing Results in Healthy and Ill Newborns: Results from the BabySeq Project. Am. J. Hum. Genet. 2019, 104, 76–93. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. The structures of human Kv7.1 (a) and human Kv7.4 (b). One of the four protomers and four bound calmodulins (one calmodulin per protomer) are shown in black and pale orange, respectively. In both (a) and (b), a single protomer is shown on the right. The residues referred to in the main text are shown in red. (c) The amino acid sequences of human Kv7.1 and human Kv7.4. The highlights indicate helices, whose colors are matched with those used in (a,b).
Figure 1. The structures of human Kv7.1 (a) and human Kv7.4 (b). One of the four protomers and four bound calmodulins (one calmodulin per protomer) are shown in black and pale orange, respectively. In both (a) and (b), a single protomer is shown on the right. The residues referred to in the main text are shown in red. (c) The amino acid sequences of human Kv7.1 and human Kv7.4. The highlights indicate helices, whose colors are matched with those used in (a,b).
Biomedicines 10 02254 g001
Table 1. JLNS-associated KCNQ1 variants that affect the amino acid sequence of the Kv7.1 protein.
Table 1. JLNS-associated KCNQ1 variants that affect the amino acid sequence of the Kv7.1 protein.
Variant *Functional Study
c.2T>Ctranslation starts
from Met159?
[21][21]
c.115G>Tp.E39X[22]
c.546C>Ap.S182R[23]
c.557G>Ap.G186D[24]
c.604G>Ap.D202N[25][26]
c.728G>Ap.R243H[27][27,28,29,30,31,32,33]
c.775C>Gp.R259G[34]
c.783G>Cp.E261D[35][29,31]
c.815G>Ap.G272D[35]
c.914G>Cp.W305S[36][27,37]
c.1040T>Gp.L347R[38]
c.1051T>Cp.F351L[24]
c.1175G>Ap.W392X[39]
c.1588C>Tp.Q530X[35][5,29,31,40,41]
c.1741A>Tp.K581X[42]
c.431delCp.I145Sfs[39]
c.443delAp.Y148Lfs[24]
c.451_452delCTp.L151Gfs[43]
c.585delGp.L196Sfs[25]
c.733_734delGGp.G245Rfs[44]
c.820_830del11p.I274Vfs[45]
c.998_999delCTp.S333Cfs[46]
c.1008delCp.I337Sfs[35][29,31]
c.1188delCp.R397Gfs[47]
c.1319delTp.V440Afs[48]
c.1356delGp.L453Wfs[49]
c.567dupGp.R190Afs[50]
c.1149dupTp.A384Cfs[21][21]
c.743_744delGGinsTCp.W248F[51][40,51]
c.1630_1635delCAGTACinsGTTGAGAp.Q544Vfs[4][27,37]
* Multiple case studies are reported for many variants, but only the primary report is provided for each. JLNS-associated KCNQ1 variants that were also reported as dominantly inherited and LQTS-associated are not included. Most functional studies reported significantly reduced or lost K+ channel activity and little to small dominant-negative inhibitory effects on Kv7.1WT, which account for the recessive inheritance of the JLNS-associated Kv7.1 variants. Two studies [30,33] reported reduced sensitivity to PIP2. Our recent study [41] examined the cell-death-inducing cytotoxicity.
Table 2. DFNA2A-associated KCNQ4 variants that affect the amino acid sequence of the Kv7.4 protein.
Table 2. DFNA2A-associated KCNQ4 variants that affect the amino acid sequence of the Kv7.4 protein.
Variant *Functional Study
c.140T>Cp.L47P[54][54]
c.343C>Gp.L115V[55]
c.463G>Ap.G155R[56]
c.546C>Gp.F182L[57][58]
c.572C>Tp.A191V[59]
c.650T>Ap.M217K[60]
c.689T>Ap.V230E[61][62]
c.701A>Tp.H234L[63]
c.709G>Ap.E237K[59]
c.725G>Ap.W242X[64][41]
c.754G>Cp.A252P[56]
c.767T>Gp.V256G[59]
c.770A>Gp.Y257C[59]
c.773T>Cp.L258P[59]
c.778G>Ap.E260K[64][62]
c.785A>Tp.D262V[64][62]
c.796G>Tp.D266Y[65][65]
c.808T>Cp.Y270H[66][62]
c.821T>Ap.L274H[67][58,68]
c.823T>Cp.W275R[69][62]
c.824G>Cp.W275S[70]
c.827G>Tp.W276L[71]
c.827G>Cp.W276S[72][58,68,73,74]
c.842T>Cp.L281S[75][58,68]
c.842T>Gp.L281W[59]
c.853G>Tp.G285C[72][58,68,73]
c.853G>Ap.G285S[76][52,68,76,77]
c.857A>Gp.Y286C[60]
c.857A>Cp.Y286S[78]
c.859G>Cp.G287R[79][62]
c.872C>Tp.P291L[61][62]
c.871C>Tp.P291S[61][62]
c.878C>Tp.T293I[59]
c.887G>Ap.G296D[80]
c.886G>Ap.G296S[81][58,68,81]
c.889A>Gp.R297G[59]
c.891G>Tp.R297S[61]
c.947G>Tp.G316V[82]
c.956G>Ap.G319D[83][83]
c.961G>Ap.G321S[72][58,68]
c.992G>Ap.R331Q[83][83]
c.992G>Cp.R331P[59]
c.1012C>Gp.R338G[84]
c.1012C>Tp.R338W[59]
c.1288G>Ap.E430K[85]
c.1316G>Ap.R439H[84]
c.1365G>Tp.H455Q[86][74,81]
c.1498C>Tp.R500C[59]
c.1600A>Gp.I534V[82]
c.1647C>Gp.F549L[87]
c.1762G>Cp.G588R[59]
c.2014G>Ap.V672M[88]
c.2039C>Tp.S680F[89][62]
c.211delCp.Q71Sfs[90][41]
c.212_224del13p.Q71Pfs[72]
c.261_269delCTACAACGTp.Y88_V90del[65][65]
c.664_681del18p.G222_L227del[73][73]
c.806_808delCCTp.S269del[91][83]
c.811_816delGCCGACp.A271_D272del[83][83]
c.1044_1051delTGCCTGGCp.A349Pfs[92][41]
c.1725delGp.I576Sfs[93]
c.228_229dupGCp.H77Rfs[61]
c.1671_1672dupACGACp.V558Tfs[94]
* Multiple case studies are reported for many variants, but only the primary report is provided for each. Seven potentially pathogenic KCNQ4 missense variants (p.N264S, p.S269F, p.S273A, p.T278A, p.L281M, p.L295P, p.R433W) reported in gnomAD and characterized in Jung et al. [74] are not included. These functional studies reported significantly reduced or lost K+ channel activity (or membrane targeting) and severe dominant-negative inhibitory effects on Kv7.4WT, except for p.F182L (Kv7.4WT-like), p.H455Q (Kv7.4WT-like), and p.G319D (nonfunctional when singly expressed, but gains hyperactivity when co-expressed with Kv7.4WT). Our recent study [41] examined the cell-death-inducing cytotoxicity. Functional characterization in a mouse model [52].
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Homma, K. The Pathological Mechanisms of Hearing Loss Caused by KCNQ1 and KCNQ4 Variants. Biomedicines 2022, 10, 2254. https://doi.org/10.3390/biomedicines10092254

AMA Style

Homma K. The Pathological Mechanisms of Hearing Loss Caused by KCNQ1 and KCNQ4 Variants. Biomedicines. 2022; 10(9):2254. https://doi.org/10.3390/biomedicines10092254

Chicago/Turabian Style

Homma, Kazuaki. 2022. "The Pathological Mechanisms of Hearing Loss Caused by KCNQ1 and KCNQ4 Variants" Biomedicines 10, no. 9: 2254. https://doi.org/10.3390/biomedicines10092254

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