Next Article in Journal
The Antioxidant Dendrobium officinale Polysaccharide Modulates Host Metabolism and Gut Microbiota to Alleviate High-Fat Diet-Induced Atherosclerosis in ApoE−/− Mice
Previous Article in Journal
Bioactive Compounds Protect Mammalian Reproductive Cells from Xenobiotics and Heat Stress-Induced Oxidative Distress via Nrf2 Signaling Activation: A Narrative Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Consolidated Understanding of the Contribution of Redox Dysregulation in the Development of Hearing Impairment

by
Xin Yi Yeo
1,2,
Soohyun Kwon
2,3,
Kimberley R. Rinai
4,
Sungsu Lee
5,
Sangyong Jung
2,* and
Raekil Park
6,*
1
Department of Psychological Medicine, Yong Loo Lin School of Medicine, National University of Singapore, Singapore 119228, Singapore
2
Department of Medical Science, College of Medicine, CHA University, Seongnam 13488, Republic of Korea
3
Department of BioNanotechnology, Gachon University, Seongnam 13120, Republic of Korea
4
Department of Life Science, College of Medicine, CHA University, Seongnam 13488, Republic of Korea
5
Department of Otolaryngology-Head and Neck Surgery, Chonnam National University Hospital and Medical School, Gwangju 61469, Republic of Korea
6
Department of Biomedical Science and Engineering, Gwangju Institute of Science & Technology (GIST), Gwangju 61005, Republic of Korea
*
Authors to whom correspondence should be addressed.
Antioxidants 2024, 13(5), 598; https://doi.org/10.3390/antiox13050598
Submission received: 22 February 2024 / Revised: 26 April 2024 / Accepted: 9 May 2024 / Published: 13 May 2024
(This article belongs to the Special Issue Oxidative Stress in Ear Damage)

Abstract

:
The etiology of hearing impairment is multifactorial, with contributions from both genetic and environmental factors. Although genetic studies have yielded valuable insights into the development and function of the auditory system, the contribution of gene products and their interaction with alternate environmental factors for the maintenance and development of auditory function requires further elaboration. In this review, we provide an overview of the current knowledge on the role of redox dysregulation as the converging factor between genetic and environmental factor-dependent development of hearing loss, with a focus on understanding the interaction of oxidative stress with the physical components of the peripheral auditory system in auditory disfunction. The potential involvement of molecular factors linked to auditory function in driving redox imbalance is an important promoter of the development of hearing loss over time.

1. Introduction

Human perception is an essential biophysical construct of the neurological system, involving the initial detection of environmental stimuli with the sensory organs [1,2,3,4], processing of the spatially dispersed information received and presented as varying local field potentials in the primary sensory cortices [5,6], and the integration of signals within the thalamocortical circuits [7,8], aiding with the identification and interpretation of environmental sensory stimuli and to comprehend the world around an individual. As such, depending on the severity of the primary sensory deficits, individuals experience limitations in physical function [9,10] and neurocognitive changes [11,12,13], which lead to societal costs in the form of economic [14] and psychosocial impacts [15,16]. Of the sensory modalities present in humans, the loss of visual and auditory perception likely has the most significant impact on the overall well-being of an individual in the visual- and audio-dominated modern society [17]. Despite the relatively well-described anatomy of the auditory system [18] and the general process of auditory signal transmission [19,20], there is a lack of understanding of the precise interactions between known environmental, physical, and biomolecular risk factors in driving auditory dysfunction beyond the limited investigation of phenotypic changes based on the previously known molecular functions of the factors.
The redox system is an important component of cellular function and homeostasis. Under physiological conditions, the generation and elimination of reactive molecular species, comprising the radical species—hydroxyl radical (HO), singlet oxygen (1O2), superoxide (O2•−), nitric oxide (NO), and peroxynitrite (ONOO)—and non-radical species such as hydrogen peroxide (H2O2) are actively balanced by the endogenous antioxidant molecules to prevent deleterious effects on the deoxyribonucleotide (DNA) [21], lipid [22], and protein structures [23]. Glutathione (GSH) is the most abundant endogenous antioxidant molecule (1–10 mM range) [24] that can directly react with O2•− and alternate antioxidant molecules to reform the antioxidant depleted during radical scavenging [25]. For instance, GSH can reduce the dehydroascorbic acid formed in the conversion of α-tocopheroxyl radical back to the α-tocopherol antioxidant. GSH is synthesized by a two-step process. The first involves the rate-limiting, adenosine triphosphate (ATP)-dependent fusion of L-cystine and L-glutamine to give γ-glutamyl-cisteine catalyzed by the γ-glutamyl-cysteine ligase (GCL), and this is followed by the GSH synthetase-dependent fusion of γ-glutamyl-cisteine with L-glycine to γ-glutamyl-cisteinyl-glicine (GSH).
In the presence of reactive oxygen species (ROS) sources, two molecules of GSH are converted to the oxidized form (GSSG) by glutathione peroxidase (GPx). GSH can be regenerated by the reduction of GSSG with glutathione reductase (GR), flavin adenine dinucleotide (FAD), and nicotinamide-adenine dinucleotide phosphate (NADPH) [26]. The dynamic conversion of GSH to GSSG, involving alternate antioxidants, and vice versa forms the key redox cycle responsible for the regulation of cellular ROS levels. GSH may alternatively react with cysteine residues in proteins, which can be reversed by glutathione S-transferase (GST), or GSSG being exported into the extracellular environment to maintain redox balance. As a further mechanism of redox control, GCL function is inhibited by excess GSH and limited by L-cysteine [27]. Alternatively, thioredoxins and glutaredoxins or alternate proteins containing redox-active disulphide bonds (CXXC) can restore protein function by reversing the oxidation of amino acid residues by ROS, where methionine residue is oxidized to sulfoxide and affects the protein structure [28,29,30]. Excessive, uncontrolled ROS production associated with the dysfunction of the electron transport chain [31], xenobiotic metabolism [32], or inflammatory events [33,34] can deplete the endogenous antioxidants, leading to the development of oxidative stress and damage to the molecular and organelle structures.
The high metabolic demand of sensory hair cells to sustain constant synaptic transmission [35,36] and the dependence of mechanotransduction and auditory function on optimal Ca2+ gradients [37,38], maintained through active uptake into the hair cells driven by the mitochondrial potential [39], lead to persistent exposure to reactive oxygen species (ROS), a normal byproduct of oxidative phosphorylation and mitochondrial respiration [40]. It is thought that dysregulation in cellular redox activity results in the net accumulation of ROS, which may lead to molecular and organelle damage [41,42] and, with persistently elevated ROS levels, activation of the apoptotic pathways, cumulating to cellular death [43]. The presence of genetic mutations that alter the direct susceptibility of the auditory system toward the deleterious effects of ROS or the age-dependent accumulation of genetic and molecular damages are suspected drivers of most forms of hearing impairment [44,45,46]. This review takes a slightly different approach from the recent reviews looking at established mechanisms of hearing loss [41,44,47,48] and attempts to build an understanding of the relationship between the known environmental and causal factors of auditory dysfunction and redox dysfunction and how the changes cumulate to hearing impairment. We will also propose potential biological nodes of the auditory process and drugs that may help in delaying or reducing hearing loss pathology.

2. Contribution of the Anatomy and Biology of the Auditory System to Its Function

The peripheral auditory system is responsible for the initial detection and encoding of the acoustic input based on the temporal and spectral factors and intensity of the auditory signals received. Hence, overall hearing perception is heavily modulated by the method of sound detection, receiving, processing, and transmission to the central nervous system (CNS) [49] (Figure 1A). The transmission of sound to the cochlea occurs initially via the physical conduction of sound waves through the outer ear structures before reaching the tympanic membrane. The transfer of sound waves occurs through the air column in the ear canal or the solid bone structure and softer tissue structures in the outer ear compartment [50,51,52] (Figure 1B). The sound wave-induced displacement of the tympanic membrane that forms the entrance to the middle ear structure leads to downstream movement of the fenestra ovalis (oval window) through a system of physical pulleys involving the malleus, incus, and stapes (ossicle) bones and the interconnected muscles and ligaments [53,54] (Figure 1C). The sealed bone structure and liquid environment of the cochlea allow the transduction of oval window displacements through the incompressible perilymph within the cochlear duct [55] to the movement of the basilar membrane and part of the organ of Corti (inner ear) [56] (Figure 1D, top). As such, the deflection of the stereocilia on the surface of hair cells physically linked to the moving tectorial membrane opens the mechanosensitive K+ channels present on the tip links [57], resulting in the influx of K+ from the high-potassium endolymph bath into and consequently depolarizing the hair cell [58]. Alterations in the volume, osmolarity, or ionic composition of the perilymph and endolymph affect the turgor pressure of the hair cells, which alters their capability to produce the electromotive force required for activation of the mechanosensitive channel [59] and synaptic vesicle release machinery, which are key determinants of hair cell and auditory function (Figure 1D). Depolarization of the hair cells activates the voltage-gated calcium channels (Cav), allowing the influx and accumulation of Ca2+ within the hair cell and triggering a Ca2+-dependent fusion of synaptic vesicles [60,61] for the glutamatergic transmission [62] of auditory signals into the CNS through the auditory nerve.
The outer and inner hair cells differ in their physical locations within the organ of Corti, the types of innervation received from auditory neurons, and their functions. The afferent neurons contact the inner hair cells (IHCs) for the transmission of sound signals into the brain, while the efferent innervations found on the outer hair cells (OHCs) likely contribute to the CNS-dependent modulation of cochlear responses [63]. In addition, a small amount of type II afferent neurons receive inputs from the OHC through en passant synapses [64]. However, these neurons are unlikely to encode significant acoustic information, as they have a low synapse activity level insufficient for neuron activation at basal conditions [65]. Physically, the OHCs mechanically amplify low-level, sound-induced membrane movement through a prestin-dependent mechanism [66], and the IHCs are the primary sensory receptors responsible for auditory signal transmission to the auditory nerve and hence the CNS [20]. The activity of the ribbon synapse is also sensitive toward the frequency, strength, and timing of the sound waves. The presence of varying amounts of large-conductance Ca2+-activated K+ (BK) channels, which allows the efflux of K+ during Cav activation, likely sharpens frequency detection and tuning by the inner hair cells within the tonotopic map of the cochlea [67], while varying the voltage dependence of the Ca2+ influx between individual ribbon synapses likely influences the population of spiral ganglion neurons (SGNs) recruited for signal propagation [68]. The resulting neurochemical signal received by the auditory nerve travels through the acoustic stria [69], superior olivary nucleus [70], and nucleus of lateral lemniscus [71] of the brain stem before passing through the inferior colliculus [72] and medial geniculate nucleus of the midbrain to arrive at the auditory cortex [73,74]. Once in the cortex, the information transmitted is further processed to influence cognition and behavioral outcomes [75,76]. Hence, auditory acuity is affected individually by the direct functional components of the auditory signal transmission process or indirectly through the cross-interactions between different components of the auditory system that lead to the modulation of signal transduction in a yet unexplored manner.

3. Established Drivers of the Loss of Auditory Function and Their Sensitivity to Dysregulation of the Redox System

Hearing loss can occur when any component of the auditory system is defective, and this is characterized into three distinct subtypes, depending on the locus of defects: conductive, sensorineural, and mixed hearing loss. Conductive hearing loss occurs when there is a mechanical disruption of sound transmission to the middle ear, often caused by a ruptured eardrum [77], cholesteatomas [78], otitis media [79], osteogenesis imperfecta [80], or any form of genetic mutation affecting the morphology or structure of the auditory canal up to the ossicle bones [81,82]. In situations without further auditory complications, conductive hearing loss is treated with surgical reconstruction [83,84] or managed with hearing aids [85,86]. Comparatively, sensorineural hearing loss (SNHL) happens with inner ear or auditory nerve damage and is directly amenable to environmental factors. Unlike conductive hearing loss, there is no effective treatment for SNHL once diagnosed [87]. The use of hearing aids is an unavoidable stopgap measure for compensating for existing hearing or hearing-related communication defects [88], while the control of lifestyle and environmental factors aims to delay the progression of hearing defects [87]. There is a significant proportion of patients that suffer from both physical and biological defects in the auditory system which manifest in the form of mixed hearing loss, including instances of auditory neuropathy spectrum disorder characterized by the absence of middle ear reflex and auditory brainstem response despite close-to-normal otoacoustic emission [89,90,91], originating from anatomical and genetic defects from birth or physical damage to the auditory system. From a different perspective, hearing impairment can occur as a temporary disruption of peripheral auditory function due to reversible changes to the passage of auditory signals to the inner ear or permanent loss of hearing capabilities due to the degeneration of hair cells and the auditory nerve. Interestingly, regeneration of auditory hair cells is possible only in the avian sensory epithelia [92], involving the F2R-like trypsin receptor 1 (F2RL1)/heparin-binding epidermal growth factor-like growth factor (HBEGF)/EGF receptor (EGFR)/extracellular signal-regulated kinase (ERK)-mediated re-entry of avian facultative stem cells into the S-phase of the cell cycle [93]. There is active yet nascent research on hair cell regeneration in mammals for the possibility of developing treatments to treat and cure hearing-related impairments [94,95]. Regardless of the types of hearing loss, genetic and environmental factors play an equal role in the loss of auditory function and perception.

3.1. Genetic Factors and Their Associated Predisposition to the Impairment of Auditory Function

The advent of DNA sequencing methods alongside classical whole-genome linkage analysis methods has allowed the systemic elucidation of the genes involved in hearing loss [96]. About half of all patients with hearing loss carry some form of genetic defect that contributes to the degeneration of auditory function [97,98], and the majority of known mutations linked to hearing impairment do not cause any other symptoms (non-syndromic) beyond auditory problems. To date, more than 200 genes are associated with hereditary hearing loss. The comprehensive and regularly updated list can be found on the Hereditary Hearing Loss Homepage (https://hereditaryhearingloss.org/, accessed on 22 February 2024). Genes linked to the development of auditory defects can be categorized based on their locus of functional impact within the auditory system, namely whether they are involved in the development and maturation of ear function or the maintenance of inner ear structural integrity for effective mechanotransduction or impact the function of the vestibular hair cells (Table 1, Figure 2).
From the preferential focus on understanding the role of identified genes in auditory function, little is known about the alternate cellular effects. An examination of the novel involvement in these genetic factors in parts of the redox pathway (Table 1, column on the speculated involvement in redox homeostasis) provides a plausible direction for future research investigating the relationship between the loss of auditory protein function, susceptibility to hearing loss, and cellular redox imbalance, particularly in the absence of cellular loss. More importantly, genetic factors with direct involvement in mitochondria function (MTCO1 cytochrome c oxidase subunit 1 and MTND1 reduced nicotinamide adenine dinucleotide (NADH)-ubiquinone oxidoreductase chain 1) [99,100], ROS production (ClpP caseinolytic mitochondrial matrix peptidase proteolytic subunit, NLRP3 nucleotide-binding domain, leucine-rich–containing family, and pyrin domain–containing-3) [101,102], development of oxidative stress (HSD17B4 17-β-hydroxysteroid dehydrogenase and WFS1 wolframin endoplasmic reticulum transmembrane glycoprotein) [103,104], and apoptosis (AIFM1 apoptosis-inducing factor and mitochondria-associated 1) [105] link activity-dependent cellular alterations with redox imbalance and pathology of the auditory tissues.
Due to the fundamental importance of mitochondria in energy production and cellular function, mutations in the mitochondrial deoxyribonucleic acid (mtDNA) of mitrochondrial function-associated genes often involve multiple organs [106]. There is a high level of heteroplasmy in the coding regions of mtDNA in non-mitotic tissues [107], and a higher level of mutational load or the presence of modifiers may be required for development of the pathology. The expression product of MT-CO1 (COX1) is the main subunit of cytochrome c oxidase complex I and the primary site of mitochondrial oxidative phosphorylation [108], while the associated NADH dehydrogenase function (protein encoded by MTND1) is essential for the oxidation of NADH, production of NAD+ for use in the citric acid cycle [109], and release of protons for maintenance of the mitochondrial membrane potential [110]. ClpP is a nuclear DNA-encoded peptidase and the catalytic subunit of the ClpXP complex involved in the degradation of misfolded or damaged proteins in the mitochondria [111]. By itself, ClpP serves as a mediator of the unfolded protein response [112,113] that is activated by mitochondrial dysfunction to restore electron transport chain function and remove excess ROS produced by the defective organelle [114]. While the ablation of ClpP in mice leads to a general upregulation of mitochondrial chaperones, accumulation of mtDNA, and increase in the expression of inflammatory markers in various organs of a ClpP null mouse [115], the association of ClpP with protease La (LON) is required for the removal of complex I and culling of ROS production in the presence of mitochondrial stress and when mitophagy fails [116]. In the context of energy-demanding auditory synaptic transmission, the deleterious effect of lower ATP production and ROS overproduction occurring with faulty ClpP and uncorrected complex I function [117] is likely exacerbated.
With the sustained presence of heightened ROS levels, the NLRP3 inflammasome may be activated through direct interaction with the thioredoxin-interacting protein (TXNIP) [118], driving cellular death. Within immune cells, a gain-of-function mutation in NLRP3 is linked to the sustained release of inflammatory factors [119,120] that can impair cochlea microcirculation [121], leading to ischemic damage. Further defects in the genetic factors involved in fatty acid metabolism (HSD17B4) [122], endoplasmic reticulum (ER) stress signaling (WFS1) [123], and caspase-independent cellular death (AIFM1) likely contribute to the increase in ease of the activation of cell death pathways and loss of functional components of the auditry system. These associations suggest the potential for a redox imbalance to disrupt normal auditory function or the contribution of hearing-specific proteins to propagation of the cellular impact and damage with initial oxidative stress induced through alternate means.
Table 1. Genes involved in hearing loss, sorted based on the locus of impact, and their speculated links to redox imbalance and cellular function.
Table 1. Genes involved in hearing loss, sorted based on the locus of impact, and their speculated links to redox imbalance and cellular function.
Gene SymbolProtein NameDFN LocusType of Hearing LossNormal FunctionSpeculated Involvement in Redox Homeostasis
Structural Defects
MYH14Myosin heavy chain 14DFNA4NS and SNon-muscle ATP-dependent molecular motors interact with cytoskeletal actin to regulate cell motility and polarity [124].miR-499 originates from the MYH14 intronic sequence and is involved in protection of cardiomyocytes [125] and neurons [126] from tissue damage-induced oxidative stress.
TECTAAlpha-tectorinDFNA8/12/21NSMajor non-collagenous structural component of the tectorial membrane [127].-
COCHCochlinDFNA9NSMajor non-collagenous component of the extracellular matrix of the inner ear. Linked to the regulation of bacteria-driven immune response in the inner ear [128,129,130].Shear stress which disrupts endothelial homeostasis and promotes oxidative stress leads to multimerization of cochlin and increased interaction with the mechanosensitive potassium channel subfamily K member 2 (TREK-1) in the ocular system [131].
MYOMyosin, class II and IIIDFNA4/11/22/48, DFNB3/30/37NS and SDevelopment and function of the cochlea duct (class II) and stereocilia of the vestibular hair cells (class III) [132,133].Myosin are differentially expressed under oxidative stress in diabetic rat brains [134].
COL11A2Type XI collagen, called the pro-alpha2(XI) chainDFNA13NSMinor fibrillar component of the tectorial membrane [135].-
CDH23Cadherin 23DFNB12NS and SCalcium-dependent cell-cell adhesion glycoprotein involved in maintaining normal organization of stereocilia bundle [136,137].Cadherin 23 regulates purine metabolism [138] involved in the modulation of cellular redox biology [139].
STRCStereocilinDFNB16NSStructural component of the stereocilia involved in the formation of horizontal top connectors of stereocilia and maintenance of the OHC bundle [140].-
TRIOBPTrio rho guanine nucleotide exchange factor and F-actin binding proteinDFNB28NSCytoskeleton-associated protein which organizes actin filaments into uniquely rootlet-like dense bundles that provide durability and rigidity to stereocilia [141].Actin is susceptible to oxidation and effects of reactive oxygen species on its functioning [142,143,144]. Specific composition of actin may be important for stereocilia function [145,146].
WHRNWhirlinDFNB31SPDZ domain-containing protein expressed at the ankle region of stereocilia. Regulates IHC stereocilia growth and differentiation and OHC stereocilia rigidity and organization during development [147,148].-
Functional Defects
GJB2/3/6Gap junction protein 2/3/6 or connexin 26/30/31DFNA2B/3A/3BNSFormation of hemichannels in the sensory epithelium, required for the formation of endolymphatic potential, which create sufficient driving force for K+ entry and depolarization of hair cells with activation of the MET channel [149].Connexin 26 ablation leads to increased oxidative stress in cochlea [45], likely through hemichannel-mediated spread of molecules that trigger redox imbalance in normal cells in the immediate periphery [150].
DIAPH1Diaphanous homolog 1 (Drosophila) proteinDFNA1SRegulate actin polymerization and microtubule dynamics to stabilize the cytoskeletal structure of hair cells [151].-
KCNQ4Kv7.4 potassium channelDFNA2ANSMaintaining cochlear ion homoeostasis and regulating hair cell membrane potential [152].-
SLC17A8Solute carrier family 17 member 8 or vesicular glutamate transporterDFNA25NSInvolved in the uptake of glutamate into the synaptic vesicles in IHCs [153].-
TMC1Transmembrane channel-like protein 1DFNB7/11NSIon-conducting pore of the MET channel complex [154,155].-
SLC26A4Solute carrier family 26 member 4 or pendrinDFNB4NSTransport negatively charged ions across the cell membrane. Involved in the function of the basal and intermediate cells of the stria vascularis to maintain the endocochlear potential [156].Pendrin knockout (KO) in mice leads to hyperpigmentation of the stria vascularis due to the increase in pH of the endolymph, which results in inhibition of cysteine uptake and glutathione synthesis by the surrounding cells [157]. Melanin synthesis is linked to oxidative stress in melanocytes [158].
TMPRSS3Transmembrane protease serine 3DFNB8NSEssential component of hair cell homeostasis and key to their survival. Precise mechanism unclear [159].-
PJVKPejvakinDFNB59NSInvolved in peroxisome proliferation in response to sound. Precise mechanism unclear [160].Pejvakin-mediated pexophagy protects auditory hair cells from noise exposure-induced oxidative stress [161].
SLC26A5PrestinDFNB62NSFunctions as the molecular motor in OHCs. Generates force of electromotility for the amplification of sound signals in OHCs [162].Oxidative stress inhibits the expression of prestin [163].
LHFPL5Lipoma high-mobility group protein gene fusion partner tetraspan subfamily member 5DFNB67NSTethers tip link to the MET
channel to establish maximal force sensitivity of the MET channel. Required for correct localization of protocadherin related 15 (PCDH15) and TMC1 to the mechanotransduction complex [164,165].
-
LOXHD1Lipoxygenase homology polycystin/lipoxygenase/alpha-toxin domains 1DFNB77NSInvolved in the mechanotranduction process in hair cells. Mechanism unknown [166,167].-
SERPINB6Serine proteinase inhibitor family B member 6DFNB91NSProtect hair cells from the leakage of lysosomal content during stress [168].Lysosomes are susceptible to oxidative stress-dependent destabilization of membrane, which leads to the release of lysosomal enzymes into the cytosol [169].
CABP2Calcium binding protein 2DFNB93NSModulator of IHC Cav1.3 function [170,171].CABP2 is a thioredoxin, which contains the redox-active dithiol/disulfide bond involved in defending against oxidative stress [172,173].
Developmental Defects
PRPS1Phosphoribosyl pyrophosphate synthetase 1DFNX1NS/SCatalyze first step of nucleotide synthesis. Involved in fetal auditory system development [174].Production of nicotinamide adenine dinucleotide (NAD) is phosphoribosyl pyrophosphate (PRPP)-dependent, and pyridine nucleotides are severely reduced in erythrocytes of patients with PRPS-1 superactivity [175].
POU3F4Pit-1/Oct-1/ Oct-2/unc-86 class 3 homeobox 4DFNX2NSInvolved in the development of the middle and inner ear [176].The related POU3F1 is degraded in the presence of oxidative stress [177].
EYA4Eyes absent transcriptional coactivator and phosphatase 4DFNA10SInvolved in embryonic auditory system development and mature inner ear function [178,179].Reduced EYA4 expression decreases single-stranded DNA accumulation following DNA damage and impairs homologous recombination [180].
GRXCR1Glutaredoxin and cysteine-rich domain containing 1DFNB25NSRequired for the morphogenesis of stereocilia in hair cells [181].-
ESRRBEstrogen-related receptor betaDFNB35SEssential for inner ear development and function [182].ERRB is a negative regulator of NF-E2-related factor 2 (Nrf2) [183], involved in the expression of detoxifying enzyme and antioxidant proteins against oxidative stress [184,185].
HGFHepatocyte growth factorDFNB39NSInvolved in the development of stria vascularis of the cochlear epithelium [186].HGF attenuates angiotensin II–induced oxidative stress in vascular smooth muscle cells [187] and protects retinal pigment epithelial cells from oxidative stress [188].
PTPRQProtein tyrosine phosphatase receptor type QDFNB84NSEssential for the maturation and function of the hair bundle in the cochlea [189].Increase in expression of the related PTPRO increases reactive oxygen species production and promotes apoptosis through the toll-like receptor 4 (TLR4)/ nuclear factor kappa light chain-enhancer of activated B cell (NF-κB) pathway [190].
NS = non-syndromic hearing loss, not associated with alternate clinical signs or symptoms; S = syndromic hearing loss, associated with alternate clinical conditions; DFNA = genetic mutation relating to hearing loss inherited through an autosomal dominant manner; DFNB = genetic mutation relating to hearing loss inherited through an autosomal recessive manner; DFNX = X-linked heritable genetic mutation related to hearing loss.

3.2. Noise-Induced Hearing Loss

Chronic exposure to excessive noise can lead to mechanical stress-induced stereocilia breakage [191], destruction of the stereocilia tip links which uncouple MET channels from the movement of the stereocilia [192], or damage the F-actin core of the stereocilia, leading to a decrease in rigidity of the stereocilia [193], which contributes to the positive feedback-driven loss of mechanotransduction and sensitivity of the inner ear toward auditory signals. Although the rapid repair of tip links is possible with the temporary recruitment of shorter PCDH15/PCDH15 tip links to replace the damaged mature PCDH15/CDH23 configuration, persistent mechanical stimulation can lead to irreversible premature hair cell death [194,195]. The functional capability of the OHC to alter their length in response to changes in sound waves made them sensitive to movement-induced alteration in cellular function and cell survival [196,197]. The transient loss and reformation of the link between stereocilia and the tectorial membrane is thought to contribute in part to the temporary auditory threshold shift (TTS) observed with noise-induced auditory trauma [198,199].
It is believed that an increase in the cochlear metabolic rate after noise exposure leads to altered mitochondria metabolism and an increase in the production of reactive oxygen species (ROS) [200]. Small amounts of ROS can perform as secondary messengers to modulate intracellular signaling. For instance, physiological levels of ROS can regulate autophagy processes through protein kinase B (AKT) [201] and adenosine monophosphate-activated protein kinase (AMPK) [202] pathways, which in turn affect the clearance of clearance of oxidized cellular components and survival of hair cells [203]. In contrast, uncontrolled generation of a large amount of ROS is linked to lipid peroxidation [204], protein oxidation [205], impairment of mitochondria [206], and cell death [207]. Furthermore, altered intracellular Ca2+ levels may further contribute to the generation of oxidative stress. The increase in free Ca2+ levels observed in the cochlea following noise exposure [196] has been linked to the activity of voltage-gated calcium channels [208,209] and amounts of extracellular Ca2+ [210]. The excess Ca2+ can intersect the Krebs cycle through the calcium-sensitive α-ketoglutarate dehydrogenase to induce the generation of superoxide and hydrogen peroxide [211].
Interestingly, noise exposure alters the microcirculation and blood flow and circulation within the cochlea [212,213], likely due to the production of 8-isoprostane-F2 alpha, a lipid peroxidation product [214,215] and a potent vasoconstrictor [216]. Following the recovery of cochlear blood flow upon the termination of a noise stimulus, the inner ear experiences an upregulation of nicotinamide-adenine dinucleotide phosphate oxidase 1 (NOX1) and dual oxidase 2 (DUOX2) expression [217], increased superoxide production [218], lipid peroxidation in the hair cells and SGNs [204], and reduced levels of the antioxidant coenzymes ubiquinone-9/10 [219] and the related mitochondria dysfunction [220], which set the stage for positive feedback-driven production of ROS. The loss of dendritic spine and peripheral axons of the bipolar sensory neurons and SGNs was observed as the immediate short-term effect of noise exposure [221,222]. Strong noise exposure is further correlated with the swelling and loss of cells from the stria vascularis [223], alongside a largely transient shift in the endocochlear potential [224]. Persistent IHC stimulation contributes to the excessive release and insufficient clearance of glutamate and overstimulation of the glutamate receptors on the postsynaptic auditory nerve terminals, which allows the massive and indiscriminate influx of ions and water from the surrounding into the postsynaptic neuron, cellular swelling, dysregulation of cellular homeostasis, and eventual cellular necrosis if uncontrolled [225,226,227].
It is possible that the hair cells have a differential tolerance toward changes in oxidative stress compared with normal cell types and between the hair cells and alternate supporting cells in the cochlea [228] due to their high level of cellular activity [229], which affects the cellular response toward changes in oxidative stress. The direct exposure of mouse hair cells to high hydrogen peroxide levels activates the p53-mediated apoptosis, delay cell cycle progression in proliferating cells, and loss of nuclear structure [230], supporting the proposition that the production of ROS or changes in oxidative stress levels may lead to direct cell demise. ROS and free radicals produced with the dysregulation of redox processes can break down the cell membrane [215], leading to disruption of the cellular structure and death. The lipid peroxidation observed in cochlea cells with exposure to noise may further propagate and broaden the impact of oxidative damage on cochlea function [231]. Sensory hair cell dysregulation cumulates to activation of the caspase 3 and 9 [232], which further trigger mitochondria dysfunction, the production of ROS, and eventual apoptotic death [233].

3.3. Exposure to Ototoxic Chemicals and Compounds

Numerous therapeutic drugs and environmental toxicants have been shown to possess ototoxic effects through disruption of the physical structure of the peripheral auditory system or the connected neurological pathways involved in auditory perception. More often than not, the clinical or physical benefit of ototoxic drug or chemical usage is higher than their potential impact on known auditory function (if any) when the drug or chemical is first developed, which contributes to their widespread and relative lack of control of use. The aminoglycoside family of antibiotics, which comprises streptomycin, kanamycin, neomycin, and gentamicin, which is used routinely to treat gram-negative bacteria infections, rapidly enters the blood-labyrinth barrier of the stria vascularis and the cochlea upon administration and persists within the endolymph and perilymph [234,235,236]. Upon uptake into the hair cells, these antibiotics can accumulate in the lysosome, disrupt the lysosome structure, and lead to the release of lysosomal content and antibiotics within the hair cells, triggering cellular degeneration [237]. Alternatively, aminoglycosides form a complex with iron [238] which interacts with polyphosphoinositides [239] and triggers the oxidation of arachidonic acid and its byproducts present on cellular membrane structures and the production of ROS [240]. Further c-Jun N-terminal kinase (JNK)-dependent recruitment of cellular components of the apoptotic pathway can drive the cells toward death [240].
Exposure toward the platinum derivatives cisplatin and carboplatin, which are commonly used in the treatment of solid tumors, is linked to the depletion of cochlea glutathione and antioxidant activities [241] as a result of the induction of superoxide production by the NADPH oxidase (NOX) 3, a member of the NADPH oxidases which is highly expressed in the inner ear and localized to the sensory epithelia [242]. The resulting indiscriminate destruction of the inner ear’s cellular structure can turn an innocuous noise signal into a noxious stimulus that further drives the decline in auditory system function [243]. Loop diuretics (ethacrynic acid, furosemide, and bumetanide) used for the treatment of high blood pressure act on sodium- and potassium-transporters and pumps within the stria vascularis to disturb the normal ionic concentration of the endolymph, leading to disruption of the ionic gradient between the endolymph and perilymph required for maintenance of the endocochlear potential and auditory transmission [244]. Although the effects of loop diuretics often go away upon the termination of use, instances of permanent hearing loss have been observed, likely due to sustained use over long durations [245] and the drug-dependent production of free radicals with ischemia reperfusion in the lateral wall (spiral ligament and stria vascularis) [246]. High doses of acetyl salicylic acid, painkillers, and alternate antipyretics (quinine and chloroquine) have potential ototoxic side effects which are not well characterized [247,248,249,250]. Toxic interactions between the aminoglycosides, cisplatin, loop diuretics, and other drugs have been observed, which contributes to the potentiation of preexisting ototoxic effects of consumed drugs [251,252,253].
Common environmental toxicants such as aromatic solvents likely affect auditory function through the formation of reactive intermediates [254] that disrupt the membrane structure of hair cells [255] or affect K+ recycling in the organ of Corti. Furthermore, the organic solvents may act on specific synaptic sites [256,257,258,259] and the acoustic reflex [260,261], which suggest functional disturbance beyond the peripheral auditory system. Exposure to phytotoxic agents commonly used as bactericides and fungicides led to the development of hearing impairment and involuntary eye movements in industrial workers [262], while the administration of germanium dioxide, a component of dry-cell batteries, electrical coils, and animal food additives, in rats and guinea pigs led to stria vascularis and supporting sensory epithelia degeneration [263]. There is also evidence from animal studies pointing to the potential ototoxicity of halogenated hydrocarbons through the effect of thyroid hormones [264,265,266,267]. Further validation (human) and mechanistic studies are required to pinpoint their true effect on auditory function.

3.4. Impact of Inflammatory Events on Auditory System Functioning and the Age-Related Decline of Auditory Function

Otitis media (OM), or middle infection, is a common form of ear infection that can occur at any age but is most frequently diagnosed in young infants up to 24 months of age [268]. Although OM can resolve spontaneously without complications [269], chronic OM can progress toward spreading the infection into the CNS (meningitis, cerebral abscess, and encephalitis), with various degrees of permanent sensorineural hearing loss [270]. The progression from acute to chronic OM has been linked to disruption of the antioxidant system [271,272], though the implication of the effect of an altered oxidative stress or antioxidant level (if any) on the cochlea function is unknown.
Mitochondria disfunction and ROS generation are dominant drivers of nucleotide-binding domain, leucine-rich-containing family, pyrin domain-containing-3 (NLRP3) inflammasome activation [273], a critical component of innate immunity [274] in part due to the transient receptor potential melastatin 2 (TRPM2)-mediated cellular calcium influx [275]. Mitochondrial-dependent autophagy may alternatively remove the dysfunctional mitochondria, putting a stop to the aberrant ROS production [276]. However, the gain of function mutation in NLRP3 and altered inflammasome activation are involved in a spectrum of autoinflammatory conditions characterized by prominent hearing disfunction [102] linked to persistent downstream activation of the resident cochlea macrophages and the release of cytokine interleukin 1 beta (IL-1β) [119]. The suppression of macrophage recruitment [277], IL-1β release [278], or alternative inflammatory mediators such as interleukin 6 (IL6) [279] attenuates the effect of the hearing function damage by environmental insults on hearing outcomes.
With time, the accumulation of cellular damage and loss of functional sensory cells from the genetic predisposition of auditory damage or exposure to an environmental ototoxic stimulus cumulates to an unavoidable decline in hearing acuity (Figure 3). Specifically, ROS production increases, while the efficacy of the antioxidant decreases with age [280], and the associated mitochondrial disfunction and oxidative stress play key roles in the development of several age-related neurological diseases [281,282]. Mice lacking superoxide dismutase 1 (SOD1), which is critical for breaking down superoxides [283], showed premature age-related loss of sensory hair cells [284], reduced thickness of the stria vascularis, and severe degeneration of the SGNs [285], while mice with inactive glutathione peroxidase 1 (GPX1) presented a significant increase in their hearing thresholds at high sound frequencies [286]. The ROS that persist within the auditory system can damage the mitochondria DNA [287], which impairs energy metabolism [288], enhances ROS generation [289], and alters mitochondria-dependent apoptotic pathways [290]. Metabolic disorders in the form of diabetes mellites, dyslipidemia, obesity, and hypertension negatively affect the development of hearing loss with age [291] via a plethora of effects on the auditory system.
As the perception of sound is a highly neurocognitive process involving numerous brain networks directly or indirectly involved in auditory processing and representation [18,292,293], the loss of primary auditory inputs results in a major restructuring of the neurological network beyond the auditory cortices [294,295]. In addition, recent studies revealed the co-occurrence of altered auditory perception in neurodegenerative conditions [296,297,298,299], with age-related etiologies. Hence, it does not come as a surprise that there are significant associations between hearing impairment and the development of dementia and alternate neurological conditions [300], particularly neuropsychiatric conditions involving multisensory perception [301,302]. Nonetheless, their causal relationship between changes in auditory perception and alternate neurological conditions has not been clearly established.

4. Targeting Redox Imbalance-Driven Hearing Impairment with Antioxidants

Physiologically, the function of signaling pathways and cellular homeostasis is intimately associated with the redox balance. The endogenous antioxidant systems within involve enzymes such as vitamins A, E, and C, flavonoids such as quercetin, and trace elements such as zinc or magnesium and proteins, which exhibit antioxidant properties [303]. This has led researchers to attempt to decrease ROS levels and alleviate the effects of various antioxidants, vitamins, and more [304].

4.1. Natural Product-Based Antioxidant Therapies

The results focused on antioxidant therapies in ROS-related hearing loss have shown the preventive effects of antioxidants based on age, noise, and ototoxicity-induced effects [305]. ROS-related hearing loss from damage to the inner ear can be prevented by reducing the generation of ROS or by enhancing the antioxidant system, especially through the administration of exogenous antioxidants, upregulating endogenous antioxidant production, and promoting an ROS scavenger system [306]. An example of an ROS scavenger is the N-acetyl L-cysteine (NAC), which has been widely studied against noise-induced oxidative stress. This is due to its ability to directly scavenge hydrogen radicals [307].
Studies have identified various agents to protect the inner ear from oxidative stress, such as vitamin A, as mentioned previously, or even flavonoids from a clean diet of fruits or known therapeutic foods, such as ginseng and coenzyme Q10 (CoQ10), as protection against or recovering from hearing loss [305]. For this reason, the use of these exogenous antioxidant systems has been commonly tested in several animal models, with promising results. Korean red ginseng not only has anti-ROS properties but also anti-apoptotic properties. Previous studies have shown that mice fed with Korean red ginseng 1 h and 1 day after noise exposure compared with those exposed 3 days later had faster recovery from hearing loss [308].

4.2. Supplemental Nutrients

Regular lacking of supplemental nutrients such as magnesium, iron, and zinc was found to increase the risk of loss of hearing [309]. In auditory neurons, when zinc is located with glutamate in the presynaptic calyces and excitatory synapses as a supplement, reports have displayed recovery from hearing loss. As zinc is an essential metal ion in physiological processes, especially as an antioxidant in Zn2+ form, it maintains the structural integrity and functions of DNA and proteins [310]. Regarding its antioxidant role, it inhibits the oxidization of NADPH, preventing the generation of ROS. Zinc also plays an important role as a cofactor of copper/zinc superoxide dismutase-1 (Cu/Zn SOD; SOD1), called the cell’s first line of antioxidant enzymes. It was reported to exist while bounded with SOD1 in rat cochlea, illustrating the elevation of the auditory brain stem (ABR) threshold and cochlear hair cell loss [311].

4.3. Ototoxic and Novel Drugs

Based on research, two types of ototoxic drug classes are used in clinical practice, which are aminoglycoside antibiotics and platinum-based anticancer drugs [312]. These have displayed damage to the hair cells in the organ of Corti through ROS production in the apoptopic pathways. Close monitoring of their potential ototoxicity and nephrotoxicity have been previously carried out [313]. Although the cells of the proximal convoluted tubules of the kidney can proliferate and recover from nephrotoxicity, the hair cells in the cochlea are not able to recuperate and recover from irreversible ototoxicity.
In addition to that, platinum-based anticancer drugs that include adenocarcinoma, squamous cell carcinoma, and undifferentiated carcinoma have toxic effects on the cochlea and neurons [314]. Cisplatin-related ototoxicity was reported to activate ROS production in the inner ear [315]. It upregulates NOX3 expression, activates its signaling pathway, and increases superoxide production in cultured cells and the cochlea [316]. Thus, disruption of the intracellular antioxidant system also contributes to both aminoglycoside- and cisplatin-induced oxidative stress [317].
However, there is still a lack of information on the fundamental mechanisms of redox homeostasis and molecular redox networks. In light of these pushbacks, there is only one clinically available drug made in 2022 that was approved by the Food and Drug Administration (FDA) of the United States, which is sodium thiosulfate. This drug acts as an antioxidant and a therapeutic agent, based on clinical trials [318,319]. Clinical tests showed successful rates of reduced cisplatin-induced ototoxicity of nearly 50% in hepatoblastoma patients. It also preserved the activity of antioxidant enzymes and alleviates hearing loss after 4–6 h of administration without interfering with the antineoplastic activity of cisplatin [320]. Nevertheless, the therapeutic effect was limited to patients under 18 years of age, while effects varied depending on the time intervals between the administration of cisplatin and sodium thiosulfate [305]. Therefore, the development of a more effective treatment for a wide range of acquired hearing loss is still crucial.

5. Conclusions

The auditory outcome depends on a myriad of factors that control the function of individual nodes for auditory signal transduction and interact with each other to maintain optimal primary auditory system functioning. Interestingly, regardless of the locus of change, the disruption of the redox system or the generation of oxidative stress serves as the mechanism for the convergence of system and cellular disturbances toward hair cell and sensory epithelial loss and overall auditory function disruption. Numerous reviews have attempted to summarize existing strategies to combat problems with the dysregulation of cellular redox homeostasis in hearing loss [44,304,321,322], which falls short of an understanding of the optimal time and duration for treatment, since overtreatment with antioxidants is potentially more problematic than beneficial [323]. It is also unclear and a general problem in the field of genomic sequencing and disease mapping how pathogenic variants of a single gene may contribute similarly or differently to a disease phenotype. Much more work is required to resolve the contribution of pathogenic mutations to environmental factors in the development of hearing loss and understand the different impact on auditory function and the relationship with the time of phenotypic appearance, how alternate non-auditory disease may delay or exacerbate the development of hearing loss, and finally how the development of an understanding of the different involvement of oxidative stress in the disruption of different auditory cell types (sensory or supporting) may educate us regarding cell-dependent vulnerability to insults and demise.

Author Contributions

Conceptualization, X.Y.Y., S.J. and R.P.; writing—original draft preparation, X.Y.Y., S.K. and S.J.; writing—review and editing, X.Y.Y., K.R.R., S.L., S.J. and R.P; critical revision of the manuscript, S.J., S.L., R.P. and X.Y.Y. All authors have read and agreed to the published version of the manuscript.

Funding

S.J. was supported by an Industry-Academic Cooperation Foundation CHA University grant (CHA-202300230001). S.L. was supported by a grant from the National Research Foundation of Korea (NRF-2021R1A2C1095057). R.P. was supported by the National Research Foundation of Korea (NRF-2018R1A5A1024340 and NRF-2022R1A2C3007463).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Martinez-Conde, S.; Macknik, S.L.; Hubel, D.H. The Role of Fixational Eye Movements in Visual Perception. Nat. Rev. Neurosci. 2004, 5, 229–240. [Google Scholar] [CrossRef]
  2. Colenbrander, A. Visual Functions and Functional Vision. Int. Congr. Ser. 2005, 1282, 482–486. [Google Scholar] [CrossRef]
  3. Sharma, A.; Kumar, R.; Aier, I.; Semwal, R.; Tyagi, P.; Varadwaj, P. Sense of Smell: Structural, Functional, Mechanistic Advancements and Challenges in Human Olfactory Research. Curr. Neuropharmacol. 2019, 17, 891–911. [Google Scholar] [CrossRef]
  4. Su, C.-Y.; Menuz, K.; Carlson, J.R. Olfactory Perception: Receptors, Cells, and Circuits. Cell 2009, 139, 45–59. [Google Scholar] [CrossRef]
  5. Pereira, A. Biophysical Mechanisms Supporting Conscious Perception: Prospects for an Artificial Astrocyte. Nat. Prec. 2011. [Google Scholar] [CrossRef]
  6. Liang, M.; Mouraux, A.; Hu, L.; Iannetti, G.D. Primary Sensory Cortices Contain Distinguishable Spatial Patterns of Activity for Each Sense. Nat. Commun. 2013, 4, 1979. [Google Scholar] [CrossRef] [PubMed]
  7. Wolff, M.; Morceau, S.; Folkard, R.; Martin-Cortecero, J.; Groh, A. A Thalamic Bridge from Sensory Perception to Cognition. Neurosci. Biobehav. Rev. 2021, 120, 222–235. [Google Scholar] [CrossRef]
  8. Tauste Campo, A.; Vázquez, Y.; Álvarez, M.; Zainos, A.; Rossi-Pool, R.; Deco, G.; Romo, R. Feed-Forward Information and Zero-Lag Synchronization in the Sensory Thalamocortical Circuit Are Modulated during Stimulus Perception. Proc. Natl. Acad. Sci. USA 2019, 116, 7513–7522. [Google Scholar] [CrossRef]
  9. Kong, H.H.; Shin, K.; Won, C.W. Association of Dual Sensory Impairment with Declining Physical Function in Community-Dwelling Older Adults. Int. J. Environ. Res. Public Health 2023, 20, 3546. [Google Scholar] [CrossRef] [PubMed]
  10. Zhou, Y.; Hu, Y.; Luo, J.; Li, Y.; Liu, H.; Sun, X.; Zhou, M. Association between Sensory Loss and Falls among Middle-Aged and Older Chinese Population: Cross-Sectional and Longitudinal Analyses. Front. Med. 2022, 8, 810159. [Google Scholar] [CrossRef]
  11. Schubert, C.R.; Cruickshanks, K.J.; Fischer, M.E.; Chen, Y.; Klein, B.E.K.; Klein, R.; Pinto, A.A. Sensory Impairments and Cognitive Function in Middle-Aged Adults. J. Gerontol. A Biol. Sci. Med. Sci. 2017, 72, 1087–1090. [Google Scholar] [CrossRef] [PubMed]
  12. Azeem, A.; Julleekeea, A.; Knight, B.; Sohail, I.; Bruyns-Haylett, M.; Sastre, M. Hearing Loss and Its Link to Cognitive Impairment and Dementia. Front. Dement. 2023, 2, 1199319. [Google Scholar] [CrossRef]
  13. Varadaraj, V.; Munoz, B.; Deal, J.A.; An, Y.; Albert, M.S.; Resnick, S.M.; Ferrucci, L.; Swenor, B.K. Association of Vision Impairment with Cognitive Decline Across Multiple Domains in Older Adults. JAMA Netw. Open 2021, 4, e2117416. [Google Scholar] [CrossRef] [PubMed]
  14. Deardorff, W.J.; Liu, P.L.; Sloane, R.; Van Houtven, C.; Pieper, C.F.; Hastings, S.N.; Cohen, H.J.; Whitson, H.E. Association of Sensory and Cognitive Impairment With Healthcare Utilization and Cost in Older Adults. J. Am. Geriatr. Soc. 2019, 67, 1617–1624. [Google Scholar] [CrossRef] [PubMed]
  15. Kiely, K.M.; Anstey, K.J.; Luszcz, M.A. Dual Sensory Loss and Depressive Symptoms: The Importance of Hearing, Daily Functioning, and Activity Engagement. Front. Hum. Neurosci. 2013, 7, 837. [Google Scholar] [CrossRef] [PubMed]
  16. Tseng, Y.-C.; Gau, B.-S.; Liu, T.-C.; Hsieh, Y.-S.; Huang, G.-S.; Lou, M.-F. Association between Sensory Impairments and Restricted Social Participation in Older Adults: A Cross-Sectional Study. Collegian 2022, 29, 850–859. [Google Scholar] [CrossRef]
  17. Hutmacher, F. What Is Our Most Important Sense? Front. Young Minds 2021, 9, 548120. [Google Scholar] [CrossRef]
  18. Peterson, D.C.; Reddy, V.; Launico, M.V.; Hamel, R.N. Neuroanatomy, Auditory Pathway. In StatPearls; StatPearls Publishing: Treasure Island, FL, USA, 2024. [Google Scholar]
  19. Aibara, R.; Welsh, J.T.; Puria, S.; Goode, R.L. Human Middle-Ear Sound Transfer Function and Cochlear Input Impedance. Hear. Res. 2001, 152, 100–109. [Google Scholar] [CrossRef]
  20. Fettiplace, R. Hair Cell Transduction, Tuning, and Synaptic Transmission in the Mammalian Cochlea. Compr. Physiol. 2017, 7, 1197–1227. [Google Scholar] [CrossRef]
  21. Dizdaroglu, M.; Jaruga, P.; Birincioglu, M.; Rodriguez, H. Free Radical-Induced Damage to DNA: Mechanisms and Measurement. Free Radic. Biol. Med. 2002, 32, 1102–1115. [Google Scholar] [CrossRef]
  22. Valgimigli, L. Lipid Peroxidation and Antioxidant Protection. Biomolecules 2023, 13, 1291. [Google Scholar] [CrossRef] [PubMed]
  23. Salvi, A.; Carrupt, P.-A.; Tillement, J.-P.; Testa, B. Structural Damage to Proteins Caused by Free Radicals: Asessment, Protection by Antioxidants, and Influence of Protein binding11Abbreviations: AAPH, 2,2′-Azobis(2-Amidinopropane) 2 HCl; and HSA, Human Serum Albumin. Biochem. Pharmacol. 2001, 61, 1237–1242. [Google Scholar] [CrossRef] [PubMed]
  24. Hwang, C.; Sinskey, A.J.; Lodish, H.F. Oxidized Redox State of Glutathione in the Endoplasmic Reticulum. Science 1992, 257, 1496–1502. [Google Scholar] [CrossRef] [PubMed]
  25. Pastore, A.; Piemonte, F. S-Glutathionylation Signaling in Cell Biology: Progress and Prospects. Eur. J. Pharm. Sci. 2012, 46, 279–292. [Google Scholar] [CrossRef] [PubMed]
  26. Espinosa-Diez, C.; Miguel, V.; Mennerich, D.; Kietzmann, T.; Sánchez-Pérez, P.; Cadenas, S.; Lamas, S. Antioxidant Responses and Cellular Adjustments to Oxidative Stress. Redox Biol. 2015, 6, 183–197. [Google Scholar] [CrossRef] [PubMed]
  27. Lu, S.C. Glutathione Synthesis. Biochim. Et Biophys. Acta (BBA)—Gen. Subj. 2013, 1830, 3143–3153. [Google Scholar] [CrossRef] [PubMed]
  28. Holmgren, A. Antioxidant Function of Thioredoxin and Glutaredoxin Systems. Antioxid. Redox Signal. 2000, 2, 811–820. [Google Scholar] [CrossRef] [PubMed]
  29. Kim, G.; Weiss, S.J.; Levine, R.L. Methionine Oxidation and Reduction in Proteins. Biochim. Biophys. Acta 2014, 1840, 901–905. [Google Scholar] [CrossRef] [PubMed]
  30. Wouters, M.A.; Fan, S.W.; Haworth, N.L. Disulfides as Redox Switches: From Molecular Mechanisms to Functional Significance. Antioxid. Redox Signal. 2010, 12, 53–91. [Google Scholar] [CrossRef]
  31. Napolitano, G.; Fasciolo, G.; Venditti, P. Mitochondrial Management of Reactive Oxygen Species. Antioxidants 2021, 10, 1824. [Google Scholar] [CrossRef]
  32. Ferrari, C.K.B. Effects of Xenobiotics on Total Antioxidant Capacity. Interdiscip. Toxicol. 2012, 5, 117–122. [Google Scholar] [CrossRef] [PubMed]
  33. Haddad, J.J.; Harb, H.L. L-γ-Glutamyl-l-Cysteinyl-Glycine (Glutathione; GSH) and GSH-Related Enzymes in the Regulation of pro- and Anti-Inflammatory Cytokines: A Signaling Transcriptional Scenario for Redox(y) Immunologic Sensor(s)? Mol. Immunol. 2005, 42, 987–1014. [Google Scholar] [CrossRef]
  34. Rahman, I.; Biswas, S.K.; Jimenez, L.A.; Torres, M.; Forman, H.J. Glutathione, Stress Responses, and Redox Signaling in Lung Inflammation. Antioxid. Redox Signal. 2005, 7, 42–59. [Google Scholar] [CrossRef] [PubMed]
  35. Puschner, B.; Schacht, J. Energy Metabolism in Cochlear Outer Hair Cells in Vitro. Hear. Res. 1997, 114, 102–106. [Google Scholar] [CrossRef]
  36. Rangaraju, V.; Calloway, N.; Ryan, T.A. Activity-Driven Local ATP Synthesis Is Required for Synaptic Function. Cell 2014, 156, 825–835. [Google Scholar] [CrossRef]
  37. Caprara, G.A.; Peng, A.W. Mechanotransduction in Mammalian Sensory Hair Cells. Mol. Cell. Neurosci. 2022, 120, 103706. [Google Scholar] [CrossRef]
  38. Nothwang, H.G.; Engel, J.; Knipper, M.; Friauf, E. L-Type Calcium Channels in the Auditory System. e-Neuroforum 2014, 5, 60–66. [Google Scholar] [CrossRef]
  39. Finkel, T.; Menazza, S.; Holmström, K.M.; Parks, R.J.; Liu, J.; Sun, J.; Liu, J.; Pan, X.; Murphy, E. The Ins and Outs of Mitochondrial Calcium. Circ. Res. 2015, 116, 1810–1819. [Google Scholar] [CrossRef] [PubMed]
  40. Murphy, M.P. How Mitochondria Produce Reactive Oxygen Species. Biochem. J. 2009, 417, 1–13. [Google Scholar] [CrossRef]
  41. Fetoni, A.R.; Paciello, F.; Rolesi, R.; Paludetti, G.; Troiani, D. Targeting Dysregulation of Redox Homeostasis in Noise-Induced Hearing Loss: Oxidative Stress and ROS Signaling. Free Radic. Biol. Med. 2019, 135, 46–59. [Google Scholar] [CrossRef]
  42. Rogers, L.K. Cellular Targets of Oxidative Stress. Curr. Opin. Toxicol. 2020, 20–21, 48–54. [Google Scholar] [CrossRef]
  43. Kannan, K.; Jain, S.K. Oxidative Stress and Apoptosis. Pathophysiology 2000, 7, 153–163. [Google Scholar] [CrossRef]
  44. Zhou, Y.; Fang, C.; Yuan, L.; Guo, M.; Xu, X.; Shao, A.; Zhang, A.; Zhou, D. Redox Homeostasis Dysregulation in Noise-Induced Hearing Loss: Oxidative Stress and Antioxidant Treatment. J. Otolaryngol—Head Neck Surg. 2023, 52, 78. [Google Scholar] [CrossRef] [PubMed]
  45. Fetoni, A.R.; Zorzi, V.; Paciello, F.; Ziraldo, G.; Peres, C.; Raspa, M.; Scavizzi, F.; Salvatore, A.M.; Crispino, G.; Tognola, G.; et al. Cx26 Partial Loss Causes Accelerated Presbycusis by Redox Imbalance and Dysregulation of Nfr2 Pathway. Redox Biol. 2018, 19, 301–317. [Google Scholar] [CrossRef] [PubMed]
  46. Jiang, H.; Talaska, A.E.; Schacht, J.; Sha, S.-H. Oxidative Imbalance in the Aging Inner Ear. Neurobiol. Aging 2007, 28, 1605–1612. [Google Scholar] [CrossRef] [PubMed]
  47. Tan, W.J.T.; Song, L. Role of Mitochondrial Dysfunction and Oxidative Stress in Sensorineural Hearing Loss. Hear. Res. 2023, 434, 108783. [Google Scholar] [CrossRef]
  48. Forouzanfar, F.; Asgharzade, S. MicroRNAs in Noise-Induced Hearing Loss and Their Regulation by Oxidative Stress and Inflammation. CDT 2020, 21, 1216–1224. [Google Scholar] [CrossRef] [PubMed]
  49. Litovsky, R. Development of the Auditory System. Handb. Clin. Neurol. 2015, 129, 55–72. [Google Scholar] [CrossRef] [PubMed]
  50. Hosoi, H.; Nishimura, T.; Shimokura, R.; Kitahara, T. Cartilage Conduction as the Third Pathway for Sound Transmission. Auris Nasus Larynx 2019, 46, 151–159. [Google Scholar] [CrossRef] [PubMed]
  51. Chan, W.X.; Yoon, Y.-J.; Kim, N. Mechanism of Bone-Conducted Hearing: Mathematical Approach. Biomech. Model. Mechanobiol. 2018, 17, 1731–1740. [Google Scholar] [CrossRef]
  52. Gan, R.Z.; Reeves, B.P.; Wang, X. Modeling of Sound Transmission from Ear Canal to Cochlea. Ann. Biomed. Eng. 2007, 35, 2180–2195. [Google Scholar] [CrossRef] [PubMed]
  53. Hüttenbrink, K.B. The Mechanics of the Middle-Ear at Static Air Pressures: The Role of the Ossicular Joints, the Function of the Middle-Ear Muscles and the Behaviour of Stapedial Prostheses. Acta Oto-Laryngol. 1988, 105, 1–35. [Google Scholar] [CrossRef] [PubMed]
  54. Hüttenbrink, K.B. The mechanics and function of the middle ear. Part 1: The ossicular chain and middle ear muscles. Laryngorhinootologie 1992, 71, 545–551. [Google Scholar] [CrossRef] [PubMed]
  55. Stenfelt, S.; Hato, N.; Goode, R.L. Fluid Volume Displacement at the Oval and Round Windows with Air and Bone Conduction Stimulation. J. Acoust. Soc. Am. 2004, 115, 797–812. [Google Scholar] [CrossRef] [PubMed]
  56. Koike, T.; Sakamoto, C.; Sakashita, T.; Hayashi, K.; Kanzaki, S.; Ogawa, K. Effects of a Perilymphatic Fistula on the Passive Vibration Response of the Basilar Membrane. Hear. Res. 2012, 283, 117–125. [Google Scholar] [CrossRef] [PubMed]
  57. Qiu, X.; Müller, U. Mechanically Gated Ion Channels in Mammalian Hair Cells. Front. Cell. Neurosci. 2018, 12, 100. [Google Scholar] [CrossRef] [PubMed]
  58. Zdebik, A.A.; Wangemann, P.; Jentsch, T.J. Potassium Ion Movement in the Inner Ear: Insights from Genetic Disease and Mouse Models. Physiology 2009, 24, 307–316. [Google Scholar] [CrossRef] [PubMed]
  59. Choi, C.-H.; Oghalai, J.S. Perilymph Osmolality Modulates Cochlear Function. Laryngoscope 2008, 118, 1621–1629. [Google Scholar] [CrossRef]
  60. Magistretti, J.; Spaiardi, P.; Johnson, S.L.; Masetto, S. Elementary Properties of Ca2+ Channels and Their Influence on Multivesicular Release and Phase-Locking at Auditory Hair Cell Ribbon Synapses. Front. Cell. Neurosci. 2015, 9, 123. [Google Scholar] [CrossRef]
  61. Krinner, S.; Butola, T.; Jung, S.; Wichmann, C.; Moser, T. RIM-Binding Protein 2 Promotes a Large Number of CaV1.3 Ca2+-Channels and Contributes to Fast Synaptic Vesicle Replenishment at Hair Cell Active Zones. Front. Cell. Neurosci. 2017, 11, 334. [Google Scholar] [CrossRef]
  62. Usami, S.; Matsubara, A.; Fujita, S.; Takumi, Y.; Ottersen, O.P. Chapter IX Glutamate Neurotransmission in the Mammalian Inner Ear. In Handbook of Chemical Neuroanatomy; Elsevier: Amsterdam, The Netherlands, 2000; Volume 18, pp. 255–271. ISBN 978-0-444-50286-5. [Google Scholar]
  63. Goutman, J.D.; Elgoyhen, A.B.; Gómez-Casati, M.E. Cochlear Hair Cells: The Sound-Sensing Machines. FEBS Lett. 2015, 589, 3354–3361. [Google Scholar] [CrossRef] [PubMed]
  64. Weisz, C.J.C.; Lehar, M.; Hiel, H.; Glowatzki, E.; Fuchs, P.A. Synaptic Transfer from Outer Hair Cells to Type II Afferent Fibers in the Rat Cochlea. J. Neurosci. 2012, 32, 9528–9536. [Google Scholar] [CrossRef] [PubMed]
  65. Weisz, C.; Glowatzki, E.; Fuchs, P. The Postsynaptic Function of Type II Cochlear Afferents. Nature 2009, 461, 1126–1129. [Google Scholar] [CrossRef]
  66. Xia, A.; Song, Y.; Wang, R.; Gao, S.S.; Clifton, W.; Raphael, P.; Chao, S.; Pereira, F.A.; Groves, A.K.; Oghalai, J.S. Prestin Regulation and Function in Residual Outer Hair Cells after Noise-Induced Hearing Loss. PLoS ONE 2013, 8, e82602. [Google Scholar] [CrossRef] [PubMed]
  67. Fettiplace, R. Diverse Mechanisms of Sound Frequency Discrimination in the Vertebrate Cochlea. Trends Neurosci. 2020, 43, 88–102. [Google Scholar] [CrossRef] [PubMed]
  68. Ohn, T.-L.; Rutherford, M.A.; Jing, Z.; Jung, S.; Duque-Afonso, C.J.; Hoch, G.; Picher, M.M.; Scharinger, A.; Strenzke, N.; Moser, T. Hair Cells Use Active Zones with Different Voltage Dependence of Ca2+ Influx to Decompose Sounds into Complementary Neural Codes. Proc. Natl. Acad. Sci. USA 2016, 113, E4716–E4725. [Google Scholar] [CrossRef]
  69. Sutherland, D.P.; Glendenning, K.K.; Masterton, R.B. Role of Acoustic Striae in Hearing: Discrimination of Sound-Source Elevation. Hear. Res. 1998, 120, 86–108. [Google Scholar] [CrossRef]
  70. Christov, F.; Nelson, E.G.; Gluth, M.B. Human Superior Olivary Nucleus Neuron Populations in Subjects with Normal Hearing and Presbycusis. Ann. Otol. Rhinol. Laryngol. 2018, 127, 527–535. [Google Scholar] [CrossRef] [PubMed]
  71. Gómez-Martínez, M.; Rincón, H.; Gómez-Álvarez, M.; Gómez-Nieto, R.; Saldaña, E. The Nuclei of the Lateral Lemniscus: Unexpected Players in the Descending Auditory Pathway. Front. Neuroanat. 2023, 17, 1242245. [Google Scholar] [CrossRef]
  72. Gruters, K.G.; Groh, J.M. Sounds and beyond: Multisensory and Other Non-Auditory Signals in the Inferior Colliculus. Front. Neural Circuits 2012, 6, 96. [Google Scholar] [CrossRef]
  73. Chen, L.; Wang, X.; Ge, S.; Xiong, Q. Medial Geniculate Body and Primary Auditory Cortex Differentially Contribute to Striatal Sound Representations. Nat. Commun. 2019, 10, 418. [Google Scholar] [CrossRef] [PubMed]
  74. Winer, J.A. The Functional Architecture of the Medial Geniculate Body and the Primary Auditory Cortex. In The Mammalian Auditory Pathway: Neuroanatomy; Webster, D.B., Popper, A.N., Fay, R.R., Eds.; Springer Handbook of Auditory Research; Springer: New York, NY, USA, 1992; Volume 1, pp. 222–409. ISBN 978-0-387-97800-0. [Google Scholar]
  75. King, A.J.; Hammond-Kenny, A.; Nodal, F.R. Multisensory Processing in the Auditory Cortex. In Multisensory Processes; Lee, A.K.C., Wallace, M.T., Coffin, A.B., Popper, A.N., Fay, R.R., Eds.; Springer Handbook of Auditory Research; Springer International Publishing: Cham, Switzerland, 2019; Volume 68, pp. 105–133. ISBN 978-3-030-10459-7. [Google Scholar]
  76. Van Bijnen, S.; Parkkonen, L.; Parviainen, T. Activity Level in Left Auditory Cortex Predicts Behavioral Performance in Inhibition Tasks in Children. NeuroImage 2022, 258, 119371. [Google Scholar] [CrossRef] [PubMed]
  77. Park, H.; Hong, S.N.; Kim, H.S.; Han, J.J.; Chung, J.; Seo, M.-W.; Oh, S.-H.; Chang, S.-O.; Lee, J.H. Determinants of Conductive Hearing Loss in Tympanic Membrane Perforation. Clin. Exp. Otorhinolaryngol. 2015, 8, 92. [Google Scholar] [CrossRef] [PubMed]
  78. Holt, J.J. Cholesteatoma and Otosclerosis: Two Slowly Progressive Causes of Hearing Loss Treatable through Corrective Surgery. Clin. Med. Res. 2003, 1, 151–154. [Google Scholar] [CrossRef] [PubMed]
  79. Liberman, M.C.; Liberman, L.D.; Maison, S.F. Chronic Conductive Hearing Loss Leads to Cochlear Degeneration. PLoS ONE 2015, 10, e0142341. [Google Scholar] [CrossRef] [PubMed]
  80. Pillion, J.P.; Vernick, D.; Shapiro, J. Hearing Loss in Osteogenesis Imperfecta: Characteristics and Treatment Considerations. Genet. Res. Int. 2011, 2011, 1–6. [Google Scholar] [CrossRef] [PubMed]
  81. Cox, T.C.; Camci, E.D.; Vora, S.; Luquetti, D.V.; Turner, E.E. The Genetics of Auricular Development and Malformation: New Findings in Model Systems Driving Future Directions for Microtia Research. Eur. J. Med. Genet. 2014, 57, 394–401. [Google Scholar] [CrossRef] [PubMed]
  82. Bartel-Friedrich, S.; Wulke, C. Classification and Diagnosis of Ear Malformations. GMS Curr. Top. Otorhinolaryngol. Head. Neck Surg. 2007, 6, Doc05. [Google Scholar]
  83. Vijayendra, H.; Parikh, B. Bone Conduction Improvement After Surgery for Conductive Hearing Loss. Indian. J. Otolaryngol. Head. Neck Surg. 2011, 63, 201–204. [Google Scholar] [CrossRef]
  84. Cheng, H.C.S.; Agrawal, S.K.; Parnes, L.S. Stapedectomy Versus Stapedotomy. Otolaryngol. Clin. N. Am. 2018, 51, 375–392. [Google Scholar] [CrossRef]
  85. Kakuki, T.; Miyata, R.; Yoshida, Y.; Kaizaki, A.; Kimura, A.; Kurashima, K.; Kuwata, R.; Takano, K. The Effects of Utilizing Cartilage Conduction Hearing Aids among Patients with Conductive Hearing Loss. Audiol. Res. 2023, 13, 408–417. [Google Scholar] [CrossRef] [PubMed]
  86. Janssen, R.M.; Hong, P.; Chadha, N.K. Bilateral Bone-Anchored Hearing Aids for Bilateral Permanent Conductive Hearing Loss: A Systematic Review. Otolaryngol.—Head Neck Surg. 2012, 147, 412–422. [Google Scholar] [CrossRef] [PubMed]
  87. Liu, F.; Han, B.; Zhou, X.; Huang, S.; Huang, J. Research Progress on the Treatment and Nursing of Sensorineural Hearing Loss. Front. Neurosci. 2023, 17, 1199946. [Google Scholar] [CrossRef] [PubMed]
  88. Jafari, Z.; Kolb, B.E.; Mohajerani, M.H. Age-Related Hearing Loss and Tinnitus, Dementia Risk, and Auditory Amplification Outcomes. Ageing Res. Rev. 2019, 56, 100963. [Google Scholar] [CrossRef] [PubMed]
  89. Valero, M.D.; Hancock, K.E.; Liberman, M.C. The Middle Ear Muscle Reflex in the Diagnosis of Cochlear Neuropathy. Hear. Res. 2016, 332, 29–38. [Google Scholar] [CrossRef] [PubMed]
  90. Starr, A.; Sininger, Y.S.; Pratt, H. The Varieties of Auditory Neuropathy. J. Basic Clin. Physiol. Pharmacol. 2000, 11, 215–230. [Google Scholar] [CrossRef] [PubMed]
  91. Starr, A.; Picton, T.W.; Sininger, Y.; Hood, L.J.; Berlin, C.I. Auditory Neuropathy. Brain 1996, 119, 741–753. [Google Scholar] [CrossRef] [PubMed]
  92. Cotanche, D.A.; Lee, K.H. Regeneration of Hair Cells in the Vestibulocochlear System of Birds and Mammals. Curr. Opin. Neurobiol. 1994, 4, 509–514. [Google Scholar] [CrossRef] [PubMed]
  93. Benkafadar, N.; Sato, M.P.; Ling, A.H.; Janesick, A.; Scheibinger, M.; Jan, T.A.; Heller, S. An Essential Signaling Cascade for Avian Auditory Hair Cell Regeneration. Dev. Cell 2024, 59, 280–291.e5. [Google Scholar] [CrossRef]
  94. Rubel, E.W.; Furrer, S.A.; Stone, J.S. A Brief History of Hair Cell Regeneration Research and Speculations on the Future. Hear. Res. 2013, 297, 42–51. [Google Scholar] [CrossRef]
  95. Xu, S.; Yang, N. Research Progress on the Mechanism of Cochlear Hair Cell Regeneration. Front. Cell. Neurosci. 2021, 15, 732507. [Google Scholar] [CrossRef]
  96. Vona, B.; Nanda, I.; Hofrichter, M.A.H.; Shehata-Dieler, W.; Haaf, T. Non-Syndromic Hearing Loss Gene Identification: A Brief History and Glimpse into the Future. Mol. Cell. Probes 2015, 29, 260–270. [Google Scholar] [CrossRef]
  97. Kochhar, A.; Hildebrand, M.S.; Smith, R.J.H. Clinical Aspects of Hereditary Hearing Loss. Genet. Med. 2007, 9, 393–408. [Google Scholar] [CrossRef]
  98. Aldè, M.; Cantarella, G.; Zanetti, D.; Pignataro, L.; La Mantia, I.; Maiolino, L.; Ferlito, S.; Di Mauro, P.; Cocuzza, S.; Lechien, J.R.; et al. Autosomal Dominant Non-Syndromic Hearing Loss (DFNA): A Comprehensive Narrative Review. Biomedicines 2023, 11, 1616. [Google Scholar] [CrossRef] [PubMed]
  99. Mkaouar-Rebai, E.; Chamkha, I.; Kammoun, T.; Alila-Fersi, O.; Aloulou, H.; Hachicha, M.; Fakhfakh, F. A Novel MT-CO1 m.6498C>A Variation Associated with the m.7444G>A Mutation in the Mitochondrial COI/tRNASer(UCN) Genes in a Patient with Hearing Impairment, Diabetes and Congenital Visual Loss. Biochem. Biophys. Res. Commun. 2013, 430, 585–591. [Google Scholar] [CrossRef]
  100. Ammar, M.; Tabebi, M.; Sfaihi, L.; Alila-Fersi, O.; Maalej, M.; Felhi, R.; Chabchoub, I.; Keskes, L.; Hachicha, M.; Fakhfakh, F.; et al. Mutational Screening in Patients with Profound Sensorineural Hearing Loss and Neurodevelopmental Delay: Description of a Novel m.3861A>C Mitochondrial Mutation in the MT-ND1 Gene. Biochem. Biophys. Res. Commun. 2016, 474, 702–708. [Google Scholar] [CrossRef] [PubMed]
  101. Jenkinson, E.M.; Rehman, A.U.; Walsh, T.; Clayton-Smith, J.; Lee, K.; Morell, R.J.; Drummond, M.C.; Khan, S.N.; Naeem, M.A.; Rauf, B.; et al. Perrault Syndrome Is Caused by Recessive Mutations in CLPP, Encoding a Mitochondrial ATP-Dependent Chambered Protease. Am. J. Hum. Genet. 2013, 92, 605–613. [Google Scholar] [CrossRef]
  102. Oziębło, D.; Leja, M.L.; Jeznach, A.; Orzechowska, M.; Skirecki, T.; Więsik-Szewczyk, E.; Furmanek, M.; Bałdyga, N.; Skarżyński, H.; Ołdak, M. Hearing Loss as the Main Clinical Presentation in NLRP3-Associated Autoinflammatory Disease. Front. Immunol. 2022, 13, 904632. [Google Scholar] [CrossRef] [PubMed]
  103. Amor, D.J.; Marsh, A.P.L.; Storey, E.; Tankard, R.; Gillies, G.; Delatycki, M.B.; Pope, K.; Bromhead, C.; Leventer, R.J.; Bahlo, M.; et al. Heterozygous Mutations in HSD17B4 Cause Juvenile Peroxisomal D-Bifunctional Protein Deficiency. Neurol. Genet. 2016, 2, e114. [Google Scholar] [CrossRef]
  104. Lim, H.D.; Lee, S.M.; Yun, Y.J.; Lee, D.H.; Lee, J.H.; Oh, S.-H.; Lee, S.-Y. WFS1 Autosomal Dominant Variants Linked with Hearing Loss: Update on Structural Analysis and Cochlear Implant Outcome. BMC Med. Genom. 2023, 16, 79. [Google Scholar] [CrossRef]
  105. Qiu, Y.; Wang, H.; Pan, H.; Guan, J.; Yan, L.; Fan, M.; Zhou, H.; Zhou, X.; Wu, K.; Jia, Z.; et al. AIFM1 Variants Associated with Auditory Neuropathy Spectrum Disorder Cause Apoptosis Due to Impaired Apoptosis-Inducing Factor Dimerization. J. Zhejiang Univ. Sci. B 2023, 24, 172–184. [Google Scholar] [CrossRef] [PubMed]
  106. Xu, Q.; Sun, P.; Feng, C.; Chen, Q.; Sun, X.; Chen, Y.; Tian, G. Varying Clinical Phenotypes of Mitochondrial DNA T12811C Mutation: A Case Series Report. Front. Med. 2022, 9, 912103. [Google Scholar] [CrossRef]
  107. Jazin, E.E.; Cavelier, L.; Eriksson, I.; Oreland, L.; Gyllensten, U. Human Brain Contains High Levels of Heteroplasmy in the Noncoding Regions of Mitochondrial DNA. Proc. Natl. Acad. Sci. USA 1996, 93, 12382–12387. [Google Scholar] [CrossRef] [PubMed]
  108. Timón-Gómez, A.; Nývltová, E.; Abriata, L.A.; Vila, A.J.; Hosler, J.; Barrientos, A. Mitochondrial Cytochrome c Oxidase Biogenesis: Recent Developments. Semin. Cell Dev. Biol. 2018, 76, 163–178. [Google Scholar] [CrossRef] [PubMed]
  109. Martínez-Reyes, I.; Chandel, N.S. Mitochondrial TCA Cycle Metabolites Control Physiology and Disease. Nat. Commun. 2020, 11, 102. [Google Scholar] [CrossRef]
  110. Alberts, B.; Johnson, A.; Lewis, J.; Raff, M.; Roberts, K.; Walter, P. Electron-Transport Chains and Their Proton Pumps. In Molecular Biology of the Cell, 4th ed.; Garland Science: New York, NY, USA, 2002. [Google Scholar]
  111. Wedam, R.; Greer, Y.E.; Wisniewski, D.J.; Weltz, S.; Kundu, M.; Voeller, D.; Lipkowitz, S. Targeting Mitochondria with ClpP Agonists as a Novel Therapeutic Opportunity in Breast Cancer. Cancers 2023, 15, 1936. [Google Scholar] [CrossRef] [PubMed]
  112. Haynes, C.M.; Petrova, K.; Benedetti, C.; Yang, Y.; Ron, D. ClpP Mediates Activation of a Mitochondrial Unfolded Protein Response in C. Elegans. Dev. Cell 2007, 13, 467–480. [Google Scholar] [CrossRef] [PubMed]
  113. Wu, G.; Xiong, Q.; Wei, X.; Wang, Y.; Hu, X.; He, G.; Liu, L.; Lai, Q.; Dai, Z.; Anushesh, D.; et al. Mitochondrial Unfolded Protein Response Gene CLPP Changes Mitochondrial Dynamics and Affects Mitochondrial Function. PeerJ 2019, 7, e7209. [Google Scholar] [CrossRef] [PubMed]
  114. Shpilka, T.; Haynes, C.M. The Mitochondrial UPR: Mechanisms, Physiological Functions and Implications in Ageing. Nat. Rev. Mol. Cell Biol. 2018, 19, 109–120. [Google Scholar] [CrossRef]
  115. Gispert, S.; Parganlija, D.; Klinkenberg, M.; Drose, S.; Wittig, I.; Mittelbronn, M.; Grzmil, P.; Koob, S.; Hamann, A.; Walter, M.; et al. Loss of Mitochondrial Peptidase Clpp Leads to Infertility, Hearing Loss plus Growth Retardation via Accumulation of CLPX, mtDNA and Inflammatory Factors. Hum. Mol. Genet. 2013, 22, 4871–4887. [Google Scholar] [CrossRef]
  116. Pryde, K.R.; Taanman, J.W.; Schapira, A.H. A LON-ClpP Proteolytic Axis Degrades Complex I to Extinguish ROS Production in Depolarized Mitochondria. Cell Rep. 2016, 17, 2522–2531. [Google Scholar] [CrossRef] [PubMed]
  117. Thelen, M.P.; Wirth, B.; Kye, M.J. Mitochondrial Defects in the Respiratory Complex I Contribute to Impaired Translational Initiation via ROS and Energy Homeostasis in SMA Motor Neurons. Acta Neuropathol. Commun. 2020, 8, 223. [Google Scholar] [CrossRef]
  118. Tschopp, J.; Schroder, K. NLRP3 Inflammasome Activation: The Convergence of Multiple Signalling Pathways on ROS Production? Nat. Rev. Immunol. 2010, 10, 210–215. [Google Scholar] [CrossRef] [PubMed]
  119. Nakanishi, H.; Kawashima, Y.; Kurima, K.; Chae, J.J.; Ross, A.M.; Pinto-Patarroyo, G.; Patel, S.K.; Muskett, J.A.; Ratay, J.S.; Chattaraj, P.; et al. NLRP3 Mutation and Cochlear Autoinflammation Cause Syndromic and Nonsyndromic Hearing Loss DFNA34 Responsive to Anakinra Therapy. Proc. Natl. Acad. Sci. USA 2017, 114, E7766–E7775. [Google Scholar] [CrossRef] [PubMed]
  120. Zhang, G.; Zheng, H.; Pyykko, I.; Zou, J. The TLR-4/NF-κB Signaling Pathway Activation in Cochlear Inflammation of Rats with Noise-Induced Hearing Loss. Hear. Res. 2019, 379, 59–68. [Google Scholar] [CrossRef] [PubMed]
  121. Scherer, E.Q.; Yang, J.; Canis, M.; Reimann, K.; Ivanov, K.; Diehl, C.D.; Backx, P.H.; Wier, W.G.; Strieth, S.; Wangemann, P.; et al. Tumor Necrosis Factor-α Enhances Microvascular Tone and Reduces Blood Flow in the Cochlea via Enhanced Sphingosine-1-Phosphate Signaling. Stroke 2010, 41, 2618–2624. [Google Scholar] [CrossRef] [PubMed]
  122. Violante, S.; Achetib, N.; Roermund, C.W.T.; Hagen, J.; Dodatko, T.; Vaz, F.M.; Waterham, H.R.; Chen, H.; Baes, M.; Yu, C.; et al. Peroxisomes Can Oxidize Medium- and Long-chain Fatty Acids through a Pathway Involving ABCD3 and HSD17B4. FASEB J. 2019, 33, 4355–4364. [Google Scholar] [CrossRef]
  123. Morikawa, S.; Blacher, L.; Onwumere, C.; Urano, F. Loss of Function of WFS1 Causes ER Stress-Mediated Inflammation in Pancreatic Beta-Cells. Front. Endocrinol. 2022, 13, 849204. [Google Scholar] [CrossRef] [PubMed]
  124. Donaudy, F.; Snoeckx, R.; Pfister, M.; Zenner, H.-P.; Blin, N.; Di Stazio, M.; Ferrara, A.; Lanzara, C.; Ficarella, R.; Declau, F.; et al. Nonmuscle Myosin Heavy-Chain Gene MYH14 Is Expressed in Cochlea and Mutated in Patients Affected by Autosomal Dominant Hearing Impairment (DFNA4). Am. J. Hum. Genet. 2004, 74, 770–776. [Google Scholar] [CrossRef] [PubMed]
  125. Wan, Q.; Xu, T.; Ding, W.; Zhang, X.; Ji, X.; Yu, T.; Yu, W.; Lin, Z.; Wang, J. miR-499-5p Attenuates Mitochondrial Fission and Cell Apoptosis via P21 in Doxorubicin Cardiotoxicity. Front. Genet. 2019, 9, 734. [Google Scholar] [CrossRef] [PubMed]
  126. Zhou, X.; Zheng, B.; Pang, L.; Che, Y.; Qi, X. Suppression of MALAT1 Alleviates Neurocyte Apoptosis and Reactive Oxygen Species Production through the miR-499-5p/SOX6 Axis in Subarachnoid Hemorrhage. J. Mol. Histol. 2022, 53, 85–96. [Google Scholar] [CrossRef] [PubMed]
  127. Verhoeven, K.; Laer, L.V.; Kirschhofer, K.; Legan, P.K.; Hughes, D.C.; Schatteman, I.; Verstreken, M.; Hauwe, P.V.; Coucke, P.; Chen, A.; et al. Mutations in the Human α-Tectorin Gene Cause Autosomal Dominant Non-Syndromic Hearing Impairment. Nat. Genet. 1998, 19, 60–62. [Google Scholar] [CrossRef] [PubMed]
  128. Robertson, N.G.; Lu, L.; Heller, S.; Merchant, S.N.; Eavey, R.D.; McKenna, M.; Nadol, J.B.; Miyamoto, R.T.; Linthicum, F.H.; Lubianca Neto, J.F.; et al. Mutations in a Novel Cochlear Gene Cause DFNA9, a Human Nonsyndromic Deafness with Vestibular Dysfunction. Nat. Genet. 1998, 20, 299–303. [Google Scholar] [CrossRef] [PubMed]
  129. Jung, J.; Yoo, J.E.; Choe, Y.H.; Park, S.C.; Lee, H.J.; Lee, H.J.; Noh, B.; Kim, S.H.; Kang, G.-Y.; Lee, K.-M.; et al. Cleaved Cochlin Sequesters Pseudomonas Aeruginosa and Activates Innate Immunity in the Inner Ear. Cell Host Microbe 2019, 25, 513–525.e6. [Google Scholar] [CrossRef]
  130. Hosokawa, S.; Mizuta, K.; Nakanishi, H.; Hashimoto, Y.; Arai, M.; Mineta, H.; Shindo, S.; Ikezono, T. Ultrastructural Localization of Cochlin in the Rat Cochlear Duct. Audiol. Neurotol. 2010, 15, 247–253. [Google Scholar] [CrossRef] [PubMed]
  131. Carreon, T.A.; Castellanos, A.; Gasull, X.; Bhattacharya, S.K. Interaction of Cochlin and Mechanosensitive Channel TREK-1 in Trabecular Meshwork Cells Influences the Regulation of Intraocular Pressure. Sci. Rep. 2017, 7, 452. [Google Scholar] [CrossRef] [PubMed]
  132. Yamamoto, N.; Okano, T.; Ma, X.; Adelstein, R.S.; Kelley, M.W. Myosin II Regulates Extension, Growth and Patterning in the Mammalian Cochlear Duct. Development 2009, 136, 1977–1986. [Google Scholar] [CrossRef]
  133. Cirilo, J.A.; Gunther, L.K.; Yengo, C.M. Functional Role of Class III Myosins in Hair Cells. Front. Cell Dev. Biol. 2021, 9, 643856. [Google Scholar] [CrossRef] [PubMed]
  134. Calábria, L.K.; Vieira da Costa, A.; da Silva Oliveira, R.J.; Ramos Deconte, S.; Nascimento, R.; de Carvalho, W.J.; de Oliveira, V.N.; Arcaro Filho, C.A.; Rezende Alves de Oliveira, L.; Goulart, L.R.; et al. Myosins Are Differentially Expressed under Oxidative Stress in Chronic Streptozotocin-Induced Diabetic Rat Brains. ISRN Neurosci. 2013, 2013, 423931. [Google Scholar] [CrossRef]
  135. Masaki, K.; Gu, J.W.; Ghaffari, R.; Chan, G.; Smith, R.J.H.; Freeman, D.M.; Aranyosi, A.J. Col11a2 Deletion Reveals the Molecular Basis for Tectorial Membrane Mechanical Anisotropy. Biophys. J. 2009, 96, 4717–4724. [Google Scholar] [CrossRef]
  136. Astuto, L.M.; Bork, J.M.; Weston, M.D.; Askew, J.W.; Fields, R.R.; Orten, D.J.; Ohliger, S.J.; Riazuddin, S.; Morell, R.J.; Khan, S.; et al. CDH23 Mutation and Phenotype Heterogeneity: A Profile of 107 Diverse Families with Usher Syndrome and Nonsyndromic Deafness. Am. J. Hum. Genet. 2002, 71, 262–275. [Google Scholar] [CrossRef] [PubMed]
  137. Jaiganesh, A.; Narui, Y.; Araya-Secchi, R.; Sotomayor, M. Beyond Cell-Cell Adhesion: Sensational Cadherins for Hearing and Balance. Cold Spring Harb. Perspect. Biol. 2018, 10, a029280. [Google Scholar] [CrossRef] [PubMed]
  138. Yang, S.; Xie, B.-L.; Dong, X.; Wang, L.; Zhu, G.; Wang, T.; Wu, W.; Lai, R.; Tao, R.; Guan, M.; et al. Cdh23 Affects Congenital Hearing Loss through Regulating Purine Metabolism. Front. Mol. Neurosci. 2023, 16, 1079529. [Google Scholar] [CrossRef] [PubMed]
  139. Savio, L.E.B.; Leite-Aguiar, R.; Alves, V.S.; Coutinho-Silva, R.; Wyse, A.T.S. Purinergic Signaling in the Modulation of Redox Biology. Redox Biol. 2021, 47, 102137. [Google Scholar] [CrossRef] [PubMed]
  140. Verpy, E.; Leibovici, M.; Michalski, N.; Goodyear, R.J.; Houdon, C.; Weil, D.; Richardson, G.P.; Petit, C. Stereocilin Connects Outer Hair Cell Stereocilia to One Another and to the Tectorial Membrane. J. Comp. Neurol. 2011, 519, 194–210. [Google Scholar] [CrossRef] [PubMed]
  141. Kitajiri, S.; Sakamoto, T.; Belyantseva, I.A.; Goodyear, R.J.; Stepanyan, R.; Fujiwara, I.; Bird, J.E.; Riazuddin, S.; Riazuddin, S.; Ahmed, Z.M.; et al. Actin-Bundling Protein TRIOBP Forms Resilient Rootlets of Hair Cell Stereocilia Essential for Hearing. Cell 2010, 141, 786–798. [Google Scholar] [CrossRef] [PubMed]
  142. Wilson, C.; Terman, J.R.; González-Billault, C.; Ahmed, G. Actin Filaments—A Target for Redox Regulation. Cytoskeleton 2016, 73, 577–595. [Google Scholar] [CrossRef] [PubMed]
  143. Milzani, A.; DalleDonne, I.; Colombo, R. Prolonged Oxidative Stress on Actin. Arch. Biochem. Biophys. 1997, 339, 267–274. [Google Scholar] [CrossRef] [PubMed]
  144. Lassing, I.; Schmitzberger, F.; Björnstedt, M.; Holmgren, A.; Nordlund, P.; Schutt, C.E.; Lindberg, U. Molecular and Structural Basis for Redox Regulation of β-Actin. J. Mol. Biol. 2007, 370, 331–348. [Google Scholar] [CrossRef]
  145. Andrade, L.R. Evidence for Changes in Beta- and Gamma-Actin Proportions during Inner Ear Hair Cell Life. Cytoskeleton 2015, 72, 282–291. [Google Scholar] [CrossRef]
  146. Patrinostro, X.; Roy, P.; Lindsay, A.; Chamberlain, C.M.; Sundby, L.J.; Starker, C.G.; Voytas, D.F.; Ervasti, J.M.; Perrin, B.J. Essential Nucleotide- and Protein-Dependent Functions of Actb/β-Actin. Proc. Natl. Acad. Sci. USA 2018, 115, 7973–7978. [Google Scholar] [CrossRef] [PubMed]
  147. Wang, H.; Du, H.; Ren, R.; Du, T.; Lin, L.; Feng, Z.; Zhao, D.; Wei, X.; Zhai, X.; Wang, H.; et al. Temporal and Spatial Assembly of Inner Ear Hair Cell Ankle Link Condensate through Phase Separation. Nat. Commun. 2023, 14, 1657. [Google Scholar] [CrossRef] [PubMed]
  148. Zou, J.; Mathur, P.D.; Zheng, T.; Wang, Y.; Almishaal, A.; Park, A.H.; Yang, J. Individual USH2 Proteins Make Distinct Contributions to the Ankle Link Complex during Development of the Mouse Cochlear Stereociliary Bundle. Hum. Mol. Genet. 2015, 24, 6944–6957. [Google Scholar] [CrossRef] [PubMed]
  149. Verselis, V.K. Connexin Hemichannels and Cochlear Function. Neurosci. Lett. 2019, 695, 40–45. [Google Scholar] [CrossRef]
  150. Ramachandran, S.; Xie, L.-H.; John, S.A.; Subramaniam, S.; Lal, R. A Novel Role for Connexin Hemichannel in Oxidative Stress and Smoking-Induced Cell Injury. PLoS ONE 2007, 2, e712. [Google Scholar] [CrossRef] [PubMed]
  151. Chiereghin, C.; Robusto, M.; Massa, V.; Castorina, P.; Ambrosetti, U.; Asselta, R.; Soldà, G. Role of Cytoskeletal Diaphanous-Related Formins in Hearing Loss. Cells 2022, 11, 1726. [Google Scholar] [CrossRef]
  152. Gao, Y.; Yechikov, S.; Vázquez, A.E.; Chen, D.; Nie, L. Impaired Surface Expression and Conductance of the KCNQ 4 Channel Lead to Sensorineural Hearing Loss. J. Cell. Mol. Medi 2013, 17, 889–900. [Google Scholar] [CrossRef] [PubMed]
  153. Qian, F.; Jiang, X.; Chai, R.; Liu, D. The Roles of Solute Carriers in Auditory Function. Front. Genet. 2022, 13, 823049. [Google Scholar] [CrossRef] [PubMed]
  154. Kurima, K.; Ebrahim, S.; Pan, B.; Sedlacek, M.; Sengupta, P.; Millis, B.A.; Cui, R.; Nakanishi, H.; Fujikawa, T.; Kawashima, Y.; et al. TMC1 and TMC2 Localize at the Site of Mechanotransduction in Mammalian Inner Ear Hair Cell Stereocilia. Cell Rep. 2015, 12, 1606–1617. [Google Scholar] [CrossRef]
  155. Fettiplace, R.; Furness, D.N.; Beurg, M. The Conductance and Organization of the TMC1-Containing Mechanotransducer Channel Complex in Auditory Hair Cells. Proc. Natl. Acad. Sci. USA 2022, 119, e2210849119. [Google Scholar] [CrossRef]
  156. Royaux, I.E.; Belyantseva, I.A.; Wu, T.; Kachar, B.; Everett, L.A.; Marcus, D.C.; Green, E.D. Localization and Functional Studies of Pendrin in the Mouse Inner Ear Provide Insight about the Etiology of Deafness in Pendred Syndrome. J. Assoc. Res. Otolaryngol. 2003, 4, 394–404. [Google Scholar] [CrossRef] [PubMed]
  157. Remigante, A.; Spinelli, S.; Pusch, M.; Sarikas, A.; Morabito, R.; Marino, A.; Dossena, S. Role of SLC4 and SLC26 Solute Carriers during Oxidative Stress. Acta Physiol. 2022, 235, e13796. [Google Scholar] [CrossRef]
  158. Kamiński, K.; Kazimierczak, U.; Kolenda, T. Oxidative Stress in Melanogenesis and Melanoma Development. Contemp. Oncol. 2022, 26, 1–7. [Google Scholar] [CrossRef] [PubMed]
  159. Tang, P.-C.; Alex, A.L.; Nie, J.; Lee, J.; Roth, A.A.; Booth, K.T.; Koehler, K.R.; Hashino, E.; Nelson, R.F. Defective Tmprss3-Associated Hair Cell Degeneration in Inner Ear Organoids. Stem Cell Rep. 2019, 13, 147–162. [Google Scholar] [CrossRef]
  160. Delmaghani, S.; Defourny, J.; Aghaie, A.; Beurg, M.; Dulon, D.; Thelen, N.; Perfettini, I.; Zelles, T.; Aller, M.; Meyer, A.; et al. Hypervulnerability to Sound Exposure through Impaired Adaptive Proliferation of Peroxisomes. Cell 2015, 163, 894–906. [Google Scholar] [CrossRef]
  161. Defourny, J.; Aghaie, A.; Perfettini, I.; Avan, P.; Delmaghani, S.; Petit, C. Pejvakin-Mediated Pexophagy Protects Auditory Hair Cells against Noise-Induced Damage. Proc. Natl. Acad. Sci. USA 2019, 116, 8010–8017. [Google Scholar] [CrossRef] [PubMed]
  162. Dallos, P.; Fakler, B. Prestin, a New Type of Motor Protein. Nat. Rev. Mol. Cell Biol. 2002, 3, 104–111. [Google Scholar] [CrossRef]
  163. Luo, X.; Xia, Y.; Li, X.-D.; Wang, J.-Y. The Effect of AP-2δ on Transcription of the Prestin Gene in HEI-OC1 Cells upon Oxidative Stress. Cell Mol. Biol. Lett. 2019, 24, 45. [Google Scholar] [CrossRef] [PubMed]
  164. Qiu, X.; Liang, X.; Llongueras, J.P.; Cunningham, C.; Müller, U. The Tetraspan LHFPL5 Is Critical to Establish Maximal Force Sensitivity of the Mechanotransduction Channel of Cochlear Hair Cells. Cell Rep. 2023, 42, 112245. [Google Scholar] [CrossRef]
  165. Erickson, T.; Pacentine, I.V.; Venuto, A.; Clemens, R.; Nicolson, T. The Lhfpl5 Ohnologs Lhfpl5a and Lhfpl5b Are Required for Mechanotransduction in Distinct Populations of Sensory Hair Cells in Zebrafish. Front. Mol. Neurosci. 2020, 12, 320. [Google Scholar] [CrossRef]
  166. Trouillet, A.; Miller, K.K.; George, S.S.; Wang, P.; Ali, N.-E.-S.; Ricci, A.; Grillet, N. Loxhd1 Mutations Cause Mechanotransduction Defects in Cochlear Hair Cells. J. Neurosci. 2021, 41, 3331–3343. [Google Scholar] [CrossRef] [PubMed]
  167. Grillet, N.; Schwander, M.; Hildebrand, M.S.; Sczaniecka, A.; Kolatkar, A.; Velasco, J.; Webster, J.A.; Kahrizi, K.; Najmabadi, H.; Kimberling, W.J.; et al. Mutations in LOXHD1, an Evolutionarily Conserved Stereociliary Protein, Disrupt Hair Cell Function in Mice and Cause Progressive Hearing Loss in Humans. Am. J. Hum. Genet. 2009, 85, 328–337. [Google Scholar] [CrossRef]
  168. Sırmacı, A.; Erbek, S.; Price, J.; Huang, M.; Duman, D.; Cengiz, F.B.; Bademci, G.; Tokgöz-Yılmaz, S.; Hişmi, B.; Özdağ, H.; et al. A Truncating Mutation in SERPINB6 Is Associated with Autosomal-Recessive Nonsyndromic Sensorineural Hearing Loss. Am. J. Hum. Genet. 2010, 86, 797–804. [Google Scholar] [CrossRef] [PubMed]
  169. Pivtoraiko, V.N.; Stone, S.L.; Roth, K.A.; Shacka, J.J. Oxidative Stress and Autophagy in the Regulation of Lysosome-Dependent Neuron Death. Antioxid. Redox Signal 2009, 11, 481–496. [Google Scholar] [CrossRef] [PubMed]
  170. Oestreicher, D.; Picher, M.M.; Rankovic, V.; Moser, T.; Pangrsic, T. Cabp2-Gene Therapy Restores Inner Hair Cell Calcium Currents and Improves Hearing in a DFNB93 Mouse Model. Front. Mol. Neurosci. 2021, 14, 689415. [Google Scholar] [CrossRef]
  171. Picher, M.M.; Gehrt, A.; Meese, S.; Ivanovic, A.; Predoehl, F.; Jung, S.; Schrauwen, I.; Dragonetti, A.G.; Colombo, R.; Van Camp, G.; et al. Ca2+ -Binding Protein 2 Inhibits Ca2+ -Channel Inactivation in Mouse Inner Hair Cells. Proc. Natl. Acad. Sci. USA 2017, 114, E1717–E1726. [Google Scholar] [CrossRef] [PubMed]
  172. Nishinaka, Y.; Masutani, H.; Nakamura, H.; Yodoi, J. Regulatory Roles of Thioredoxin in Oxidative Stress-Induced Cellular Responses. Redox Rep. 2001, 6, 289–295. [Google Scholar] [CrossRef]
  173. Lu, J.; Holmgren, A. The Thioredoxin Antioxidant System. Free Radic. Biol. Med. 2014, 66, 75–87. [Google Scholar] [CrossRef]
  174. Mittal, R.; Patel, K.; Mittal, J.; Chan, B.; Yan, D.; Grati, M.; Liu, X.Z. Association of PRPS1 Mutations with Disease Phenotypes. Dis. Markers 2015, 2015, 1–7. [Google Scholar] [CrossRef]
  175. Simmonds, H.; Webster, D.; Lingam, S.; Wilson, J. An Inborn Error of Purine Metabolism, Deafness and Neurodevelopmental Abnormality. Neuropediatrics 1985, 16, 106–108. [Google Scholar] [CrossRef]
  176. Bernardinelli, E.; Roesch, S.; Simoni, E.; Marino, A.; Rasp, G.; Astolfi, L.; Sarikas, A.; Dossena, S. Novel POU3F4 Variants Identified in Patients with Inner Ear Malformations Exhibit Aberrant Cellular Distribution and Lack of SLC6A20 Transcriptional Upregulation. Front. Mol. Neurosci. 2022, 15, 999833. [Google Scholar] [CrossRef] [PubMed]
  177. Li, G.; Jiapaer, Z.; Weng, R.; Hui, Y.; Jia, W.; Xi, J.; Wang, G.; Zhu, S.; Zhang, X.; Feng, D.; et al. Dysregulation of the SIRT1/OCT6 Axis Contributes to Environmental Stress-Induced Neural Induction Defects. Stem Cell Rep. 2017, 8, 1270–1286. [Google Scholar] [CrossRef] [PubMed]
  178. Wayne, S. Mutations in the Transcriptional Activator EYA4 Cause Late-Onset Deafness at the DFNA10 Locus. Hum. Mol. Genet. 2001, 10, 195–200. [Google Scholar] [CrossRef]
  179. Wang, L.; Sewell, W.F.; Kim, S.D.; Shin, J.T.; MacRae, C.A.; Zon, L.I.; Seidman, J.G.; Seidman, C.E. Eya4 Regulation of Na+/K+-ATPase Is Required for Sensory System Development in Zebrafish. Development 2008, 135, 3425–3434. [Google Scholar] [CrossRef] [PubMed]
  180. de la Peña Avalos, B.; Paquet, N.; Tropée, R.; Coulombe, Y.; Palacios, H.; Leung, J.W.; Masson, J.-Y.; Duijf, P.H.G.; Dray, E. The Protein Phosphatase EYA4 Promotes Homologous Recombination (HR) through Dephosphorylation of Tyrosine 315 on RAD51. Nucleic Acids Res. 2024, 52, 1173–1187. [Google Scholar] [CrossRef]
  181. Liu, C.; Zhao, B. Murine GRXCR1 Has a Different Function Than GRXCR2 in the Morphogenesis of Stereocilia. Front. Cell. Neurosci. 2021, 15, 714070. [Google Scholar] [CrossRef]
  182. Collin, R.W.J.; Kalay, E.; Tariq, M.; Peters, T.; Van Der Zwaag, B.; Venselaar, H.; Oostrik, J.; Lee, K.; Ahmed, Z.M.; Çaylan, R.; et al. Mutations of ESRRB Encoding Estrogen-Related Receptor Beta Cause Autosomal-Recessive Nonsyndromic Hearing Impairment DFNB35. Am. J. Hum. Genet. 2008, 82, 125–138. [Google Scholar] [CrossRef]
  183. Zhou, W.; Lo, S.-C.; Liu, J.-H.; Hannink, M.; Lubahn, D.B. ERRβ: A Potent Inhibitor of Nrf2 Transcriptional Activity. Mol. Cell. Endocrinol. 2007, 278, 52–62. [Google Scholar] [CrossRef]
  184. Itoh, K.; Chiba, T.; Takahashi, S.; Ishii, T.; Igarashi, K.; Katoh, Y.; Oyake, T.; Hayashi, N.; Satoh, K.; Hatayama, I.; et al. An Nrf2/Small Maf Heterodimer Mediates the Induction of Phase II Detoxifying Enzyme Genes through Antioxidant Response Elements. Biochem. Biophys. Res. Commun. 1997, 236, 313–322. [Google Scholar] [CrossRef]
  185. Chanas, S.A.; Jiang, Q.; McMAHON, M.; McWALTER, G.K.; McLELLAN, L.I.; Elcombe, C.R.; Henderson, C.J.; Wolf, C.R.; Moffat, G.J.; Itoh, K.; et al. Loss of the Nrf2 Transcription Factor Causes a Marked Reduction in Constitutive and Inducible Expression of the Glutathione S-Transferase Gsta1, Gsta2, Gstm1, Gstm2, Gstm3 and Gstm4 Genes in the Livers of Male and Female Mice. Biochem. J. 2002, 365, 405–416. [Google Scholar] [CrossRef]
  186. Shibata, S.; Miwa, T.; Wu, H.-H.; Levitt, P.; Ohyama, T. Hepatocyte Growth Factor-c-MET Signaling Mediates the Development of Nonsensory Structures of the Mammalian Cochlea and Hearing. J. Neurosci. 2016, 36, 8200–8209. [Google Scholar] [CrossRef]
  187. Shimizu, K.; Taniyama, Y.; Sanada, F.; Azuma, J.; Iwabayashi, M.; Iekushi, K.; Rakugi, H.; Morishita, R. Hepatocyte Growth Factor Inhibits Lipopolysaccharide-Induced Oxidative Stress via Epithelial Growth Factor Receptor Degradation. ATVB 2012, 32, 2687–2693. [Google Scholar] [CrossRef] [PubMed]
  188. Jin, M.; Yaung, J.; Kannan, R.; He, S.; Ryan, S.J.; Hinton, D.R. Hepatocyte Growth Factor Protects RPE Cells from Apoptosis Induced by Glutathione Depletion. Investig. Ophthalmol. Vis. Sci. 2005, 46, 4311. [Google Scholar] [CrossRef] [PubMed]
  189. Goodyear, R.J.; Jones, S.M.; Sharifi, L.; Forge, A.; Richardson, G.P. Hair Bundle Defects and Loss of Function in the Vestibular End Organs of Mice Lacking the Receptor-Like Inositol Lipid Phosphatase PTPRQ. J. Neurosci. 2012, 32, 2762–2772. [Google Scholar] [CrossRef] [PubMed]
  190. Liang, C.; Wang, X.; Hu, J.; Lian, X.; Zhu, T.; Zhang, H.; Gu, N.; Li, J. PTPRO Promotes Oxidized Low-Density Lipoprotein Induced Oxidative Stress and Cell Apoptosis through Toll-Like Receptor 4/Nuclear Factor κB Pathway. Cell Physiol. Biochem. 2017, 42, 495–505. [Google Scholar] [CrossRef] [PubMed]
  191. Wagner, E.L.; Shin, J.-B. Mechanisms of Hair Cell Damage and Repair. Trends Neurosci. 2019, 42, 414–424. [Google Scholar] [CrossRef] [PubMed]
  192. Assad, J.A.; Shepherd, G.M.G.; Corey, D.P. Tip-Link Integrity and Mechanical Transduction in Vertebrate Hair Cells. Neuron 1991, 7, 985–994. [Google Scholar] [CrossRef]
  193. Liberman, M.C. Chronic Ultrastructural Changes in Acoustic Trauma: Serial-Section Reconstruction of Stereocilia and Cuticular Plates. Hear. Res. 1987, 26, 65–88. [Google Scholar] [CrossRef] [PubMed]
  194. Jia, S.; Yang, S.; Guo, W.; He, D.Z.Z. Fate of Mammalian Cochlear Hair Cells and Stereocilia after Loss of the Stereocilia. J. Neurosci. 2009, 29, 15277–15285. [Google Scholar] [CrossRef]
  195. Indzhykulian, A.A.; Stepanyan, R.; Nelina, A.; Spinelli, K.J.; Ahmed, Z.M.; Belyantseva, I.A.; Friedman, T.B.; Barr-Gillespie, P.G.; Frolenkov, G.I. Molecular Remodeling of Tip Links Underlies Mechanosensory Regeneration in Auditory Hair Cells. PLoS Biol. 2013, 11, e1001583. [Google Scholar] [CrossRef]
  196. Fridberger, A.; Flock, A.; Ulfendahl, M.; Flock, B. Acoustic Overstimulation Increases Outer Hair Cell Ca2+ Concentrations and Causes Dynamic Contractions of the Hearing Organ. Proc. Natl. Acad. Sci. USA 1998, 95, 7127–7132. [Google Scholar] [CrossRef]
  197. Waqas, M.; Gao, S.; Ali, M.K.; Ma, Y.; Li, W. Inner Ear Hair Cell Protection in Mammals against the Noise-Induced Cochlear Damage. Neural Plast. 2018, 2018, 1–9. [Google Scholar] [CrossRef]
  198. Nordmann, A.S.; Bohne, B.A.; Harding, G.W. Histopathological Differences between Temporary and Permanent Threshold Shift. Hear. Res. 2000, 139, 13–30. [Google Scholar] [CrossRef]
  199. Saunders, J.C.; Flock, Å. Recovery of Threshold Shift in Hair-Cell Stereocilia Following Exposure to Intense Stimulation. Hear. Res. 1986, 23, 233–243. [Google Scholar] [CrossRef]
  200. Henderson, D.; Bielefeld, E.C.; Harris, K.C.; Hu, B.H. The Role of Oxidative Stress in Noise-Induced Hearing Loss. Ear Hear. 2006, 27, 1–19. [Google Scholar] [CrossRef]
  201. Kma, L.; Baruah, T.J. The Interplay of ROS and the PI3K/Akt Pathway in Autophagy Regulation. Biotech. App Biochem. 2022, 69, 248–264. [Google Scholar] [CrossRef]
  202. Agostini, F.; Bisaglia, M.; Plotegher, N. Linking ROS Levels to Autophagy: The Key Role of AMPK. Antioxidants 2023, 12, 1406. [Google Scholar] [CrossRef]
  203. Liu, C.; Zheng, Z.; Wang, P.; He, S.; He, Y. Autophagy: A Novel Horizon for Hair Cell Protection. Neural Plast. 2021, 2021, 5511010. [Google Scholar] [CrossRef]
  204. Ohinata, Y.; Miller, J.M.; Schacht, J. Protection from Noise-Induced Lipid Peroxidation and Hair Cell Loss in the Cochlea. Brain Res. 2003, 966, 265–273. [Google Scholar] [CrossRef]
  205. Rouyère, C.; Serrano, T.; Frémont, S.; Echard, A. Oxidation and Reduction of Actin: Origin, Impact in Vitro and Functional Consequences in Vivo. Eur. J. Cell Biol. 2022, 101, 151249. [Google Scholar] [CrossRef]
  206. Palma, F.R.; Gantner, B.N.; Sakiyama, M.J.; Kayzuka, C.; Shukla, S.; Lacchini, R.; Cunniff, B.; Bonini, M.G. ROS Production by Mitochondria: Function or Dysfunction? Oncogene 2024, 43, 295–303. [Google Scholar] [CrossRef]
  207. Redza-Dutordoir, M.; Averill-Bates, D.A. Activation of Apoptosis Signalling Pathways by Reactive Oxygen Species. Biochim. Et Biophys. Acta (BBA)—Mol. Cell Res. 2016, 1863, 2977–2992. [Google Scholar] [CrossRef]
  208. Heinrich, U.-R.; Maurer, J.; Mann, W. Ultrastructural Evidence for Protection of the Outer Hair Cells of the Inner Ear during Intense Noise Exposure by Application of the Organic Calcium Channel Blocker Diltiazem. ORL 1999, 61, 321–327. [Google Scholar] [CrossRef]
  209. Shen, H.; Zhang, B.; Shin, J.-H.; Lei, D.; Du, Y.; Gao, X.; Wang, Q.; Ohlemiller, K.K.; Piccirillo, J.; Bao, J. Prophylactic and Therapeutic Functions of T-Type Calcium Blockers against Noise-Induced Hearing Loss. Hear. Res. 2007, 226, 52–60. [Google Scholar] [CrossRef]
  210. Orrenius, S.; Zhivotovsky, B.; Nicotera, P. Regulation of Cell Death: The Calcium-Apoptosis Link. Nat. Rev. Mol. Cell Biol. 2003, 4, 552–565. [Google Scholar] [CrossRef]
  211. Tretter, L.; Adam-Vizi, V. Generation of Reactive Oxygen Species in the Reaction Catalyzed by α-Ketoglutarate Dehydrogenase. J. Neurosci. 2004, 24, 7771–7778. [Google Scholar] [CrossRef]
  212. Lamm, K.; Arnold, W. Noise-Induced Cochlear Hypoxia Is Intensity Dependent, Correlates with Hearing Loss and Precedes Reduction of Cochlear Blood Flow. Audiol. Neurootol. 1996, 1, 148–160. [Google Scholar] [CrossRef]
  213. Arpornchayanon, W.; Canis, M.; Suckfuell, M.; Ihler, F.; Olzowy, B.; Strieth, S. Modeling the Measurements of Cochlear Microcirculation and Hearing Function after Loud Noise. Otolaryngol. Head. Neck Surg. 2011, 145, 463–469. [Google Scholar] [CrossRef]
  214. Marley, R.; Harry, D.; Anand, R.; Fernando, B.; Davies, S.; Moore, K. 8-Isoprostaglandin F2 Alpha, a Product of Lipid Peroxidation, Increases Portal Pressure in Normal and Cirrhotic Rats. Gastroenterology 1997, 112, 208–213. [Google Scholar] [CrossRef]
  215. Ohinata, Y.; Miller, J.M.; Altschuler, R.A.; Schacht, J. Intense Noise Induces Formation of Vasoactive Lipid Peroxidation Products in the Cochlea. Brain Res. 2000, 878, 163–173. [Google Scholar] [CrossRef]
  216. Miller, J.M.; Brown, J.N.; Schacht, J. 8-Iso-Prostaglandin F, a Product of Noise Exposure, Reduces Inner Ear Blood Flow. Audiol. Neurotol. 2003, 8, 207–221. [Google Scholar] [CrossRef]
  217. Vlajkovic, S.M.; Lin, S.C.; Wong, A.C.Y.; Wackrow, B.; Thorne, P.R. Noise-Induced Changes in Expression Levels of NADPH Oxidases in the Cochlea. Hear. Res. 2013, 304, 145–152. [Google Scholar] [CrossRef]
  218. Yamane, H.; Nakai, Y.; Takayama, M.; Iguchi, H.; Nakagawa, T.; Kojima, A. Appearance of Free Radicals in the Guinea Pig Inner Ear after Noise-Induced Acoustic Trauma. Eur. Arch. Otorhinolaryngol. 1995, 252, 504–508. [Google Scholar] [CrossRef]
  219. Fetoni, A.R.; De Bartolo, P.; Eramo, S.L.M.; Rolesi, R.; Paciello, F.; Bergamini, C.; Fato, R.; Paludetti, G.; Petrosini, L.; Troiani, D. Noise-Induced Hearing Loss (NIHL) as a Target of Oxidative Stress-Mediated Damage: Cochlear and Cortical Responses after an Increase in Antioxidant Defense. J. Neurosci. 2013, 33, 4011–4023. [Google Scholar] [CrossRef]
  220. Yang, Z.-J.; Zhao, C.-L.; Liang, W.-Q.; Chen, Z.-R.; Du, Z.-D.; Gong, S.-S. ROS-Induced Oxidative Stress and Mitochondrial Dysfunction: A Possible Mechanism Responsible for Noise-Induced Ribbon Synaptic Damage. Am. J. Transl. Res. 2024, 16, 272–284. [Google Scholar] [CrossRef]
  221. Kujawa, S.G.; Liberman, M.C. Adding Insult to Injury: Cochlear Nerve Degeneration after “Temporary” Noise-Induced Hearing Loss. J. Neurosci. 2009, 29, 14077–14085. [Google Scholar] [CrossRef]
  222. Yan, W.; Liu, W.; Qi, J.; Fang, Q.; Fan, Z.; Sun, G.; Han, Y.; Zhang, D.; Xu, L.; Wang, M.; et al. A Three-Dimensional Culture System with Matrigel Promotes Purified Spiral Ganglion Neuron Survival and Function In Vitro. Mol. Neurobiol. 2018, 55, 2070–2084. [Google Scholar] [CrossRef]
  223. Wang, Y.; Hirose, K.; Liberman, M.C. Dynamics of Noise-Induced Cellular Injury and Repair in the Mouse Cochlea. J. Assoc. Res. Otolaryngol. 2002, 3, 248–268. [Google Scholar] [CrossRef]
  224. Hirose, K.; Liberman, M.C. Lateral Wall Histopathology and Endocochlear Potential in the Noise-Damaged Mouse Cochlea. JARO—J. Assoc. Res. Otolaryngol. 2003, 4, 339–352. [Google Scholar] [CrossRef]
  225. Spoendlin, H. Primary Structural Changes in the Organ of Corti After Acoustic Overstimulation. Acta Oto-Laryngol. 1971, 71, 166–176. [Google Scholar] [CrossRef]
  226. Puel, J.L.; Ruel, J.; Gervais d’Aldin, C.; Pujol, R. Excitotoxicity and Repair of Cochlear Synapses after Noise-Trauma Induced Hearing Loss. Neuroreport 1998, 9, 2109–2114. [Google Scholar] [CrossRef]
  227. Pujol, R.; Rebillard, G.; Puel, J.-L.; Lenoir, M.; Eybalin, M.; Recasens, M. Glutamate Neurotoxicity in the Cochlea: A Possible Consequence of Ischaemic or Anoxic Conditions Occurring in Ageing. Acta Oto-Laryngol. 1991, 111, 32–36. [Google Scholar] [CrossRef]
  228. Baker, K.; Staecker, H. Low Dose Oxidative Stress Induces Mitochondrial Damage in Hair Cells. Anat. Rec. 2012, 295, 1868–1876. [Google Scholar] [CrossRef]
  229. Bozovic, D. Active Biomechanics of Sensory Hair Bundles. Cold Spring Harb. Perspect. Med. 2019, 9, a035014. [Google Scholar] [CrossRef] [PubMed]
  230. Gentilin, E.; Cani, A.; Simoni, E.; Chicca, M.; Di Paolo, M.L.; Martini, A.; Astolfi, L. Hydrogen Peroxide Toxicity on Auditory Cells: An in Vitro Study. Chem.-Biol. Interact. 2021, 345, 109575. [Google Scholar] [CrossRef]
  231. Ayala, A.; Muñoz, M.F.; Argüelles, S. Lipid Peroxidation: Production, Metabolism, and Signaling Mechanisms of Malondialdehyde and 4-Hydroxy-2-Nonenal. Oxid. Med. Cell Longev. 2014, 2014, 360438. [Google Scholar] [CrossRef]
  232. Cheng, A.G.; Cunningham, L.L.; Rubel, E.W. Mechanisms of Hair Cell Death and Protection. Curr. Opin. Otolaryngol. Head Neck Surg. 2005, 13, 343–348. [Google Scholar] [CrossRef] [PubMed]
  233. Brentnall, M.; Rodriguez-Menocal, L.; De Guevara, R.L.; Cepero, E.; Boise, L.H. Caspase-9, Caspase-3 and Caspase-7 Have Distinct Roles during Intrinsic Apoptosis. BMC Cell Biol. 2013, 14, 32. [Google Scholar] [CrossRef] [PubMed]
  234. Tran Ba Huy, P.; Meulemans, A.; Wassef, M.; Manuel, C.; Sterkers, O.; Amiel, C. Gentamicin Persistence in Rat Endolymph and Perilymph after a Two-Day Constant Infusion. Antimicrob. Agents Chemother. 1983, 23, 344–346. [Google Scholar] [CrossRef]
  235. Dulon, D.; Hiel, H.; Aurousseau, C.; Erre, J.P.; Aran, J.M. Pharmacokinetics of Gentamicin in the Sensory Hair Cells of the Organ of Corti: Rapid Uptake and Long Term Persistence. C R. Acad. Sci. III 1993, 316, 682–687. [Google Scholar]
  236. Aran, J.; Erre, J.; Da Costa, D.L.; Debbarh, I.; Dulon, D. Acute and Chronic Effects of Aminoglycosides on Cochlear Hair Cells. Ann. N. Y. Acad. Sci. 1999, 884, 60–68. [Google Scholar] [CrossRef] [PubMed]
  237. Hashino, E.; Shero, M.; Salvi, R.J. Lysosomal Targeting and Accumulation of Aminoglycoside Antibiotics in Sensory Hair Cells. Brain Res. 1997, 777, 75–85. [Google Scholar] [CrossRef] [PubMed]
  238. Priuska, E.M.; Schacht, J. Formation of Free Radicals by Gentamicin and Iron and Evidence for an Iron/Gentamicin Complex. Biochem. Pharmacol. 1995, 50, 1749–1752. [Google Scholar] [CrossRef] [PubMed]
  239. Xie, J.; Talaska, A.E.; Schacht, J. New Developments in Aminoglycoside Therapy and Ototoxicity. Hear. Res. 2011, 281, 28–37. [Google Scholar] [CrossRef]
  240. Karasawa, T.; Steyger, P.S. Intracellular Mechanisms of Aminoglycoside-Induced Cytotoxicity. Integr. Biol. 2011, 3, 879–886. [Google Scholar] [CrossRef] [PubMed]
  241. Ravi, R.; Somani, S.M.; Rybak, L.P. Mechanism of Cisplatin Ototoxicity: Antioxidant System. Pharmacol. Toxicol. 1995, 76, 386–394. [Google Scholar] [CrossRef] [PubMed]
  242. Bánfi, B.; Malgrange, B.; Knisz, J.; Steger, K.; Dubois-Dauphin, M.; Krause, K.-H. NOX3, a Superoxide-Generating NADPH Oxidase of the Inner Ear. J. Biol. Chem. 2004, 279, 46065–46072. [Google Scholar] [CrossRef] [PubMed]
  243. Don Brown, R.; Penny, J.E.; Henley, C.M.; Hodges, K.B.; Kupetz, S.A.; Glenn, D.W.; Jobe, P.C. Ototoxic Drugs and Noise. In Novartis Foundation Symposia; Evered, D., Lawrenson, G., Eds.; Wiley: Hoboken, NJ, USA, 1981; Volume 85, pp. 151–171. ISBN 978-0-470-66395-0. [Google Scholar]
  244. Forge, A. A Tubulo-Cisternal Endoplasmic Reticulum System in the Potassium Transporting Marginal Cells of the Stria Vascularis and Effects of the Ototoxic Diuretic Ethacrynic Acid. Cell Tissue Res. 1982, 226, 375–387. [Google Scholar] [CrossRef] [PubMed]
  245. Martínez-Rodríguez, R.; Bellido Peti, J.; Palou Redorta, J.; Gómez Ruiz, J.J.; Villavicencio Mavrich, H.; García Lorenzo, J. Diuréticos del asa y ototoxicidad. Actas Urológicas Españolas 2007, 31, 1189–1192. [Google Scholar] [CrossRef]
  246. Ding, D.; McFadden, S.L.; Woo, J.M.; Salvi, R.J. Ethacrynic Acid Rapidly and Selectively Abolishes Blood Flow in Vessels Supplying the Lateral Wall of the Cochlea. Hear. Res. 2002, 173, 1–9. [Google Scholar] [CrossRef]
  247. McCabe, P.A.; Dey, F.L. XXVIII The Effect of Aspirin upon Auditory Sensitivity. Ann. Otol. Rhinol. Laryngol. 1965, 74, 312–325. [Google Scholar] [CrossRef] [PubMed]
  248. Johnson, A.C. The Ototoxic Effect of Toluene and the Influence of Noise, Acetyl Salicylic Acid, or Genotype. A Study in Rats and Mice. Scand. Audiol. Suppl. 1993, 39, 1–40. [Google Scholar] [PubMed]
  249. Curhan, S.G.; Eavey, R.; Shargorodsky, J.; Curhan, G.C. Analgesic Use and the Risk of Hearing Loss in Men. Am. J. Med. 2010, 123, 231–237. [Google Scholar] [CrossRef] [PubMed]
  250. Fernandes, C.P.; Vernier, L.S.; Dallegrave, E.; Machado, M.S. The Ototoxicity of Chloroquine and Hydroxychloroquine: A Systematic Review. Int. Arch. Otorhinolaryngol. 2022, 26, e167–e177. [Google Scholar] [CrossRef]
  251. Walker, E.M.; Fazekas-May, M.A.; Bowen, W.R. Nephrotoxic and Ototoxic Agents. Clin. Lab. Med. 1990, 10, 323–354. [Google Scholar] [CrossRef] [PubMed]
  252. Komune, S.; Snow, J.B. Potentiating Effects of Cisplatin and Ethacrynic Acid in Ototoxicity. Arch. Otolaryngol.—Head Neck Surg. 1981, 107, 594–597. [Google Scholar] [CrossRef] [PubMed]
  253. Whitworth, C.A.; Hudson, T.E.; Rybak, L.P. The Effect of Combined Administration of Cadmium and Furosemide on Auditory Function in the Rat. Hear. Res. 1999, 129, 61–70. [Google Scholar] [CrossRef] [PubMed]
  254. Chen, G.-D.; Chi, L.-H.; Kostyniak, P.J.; Henderson, D. Styrene Induced Alterations in Biomarkers of Exposure and Effects in the Cochlea: Mechanisms of Hearing Loss. Toxicol. Sci. 2007, 98, 167–177. [Google Scholar] [CrossRef]
  255. Campo, P.; Lataye, R.; Loquet, G.; Bonnet, P. Styrene-Induced Hearing Loss: A Membrane Insult. Hear. Res. 2001, 154, 170–180. [Google Scholar] [CrossRef]
  256. Cruz, S.L.; Mirshahi, T.; Thomas, B.; Balster, R.L.; Woodward, J.J. Effects of the Abused Solvent Toluene on Recombinant N-Methyl-D-Aspartate and Non-N-Methyl-D-Aspartate Receptors Expressed in Xenopus Oocytes. J. Pharmacol. Exp. Ther. 1998, 286, 334–340. [Google Scholar]
  257. Beckstead, M.J.; Weiner, J.L.; Eger, E.I.; Gong, D.H.; Mihic, S.J. Glycine and Gamma-Aminobutyric Acid(A) Receptor Function Is Enhanced by Inhaled Drugs of Abuse. Mol. Pharmacol. 2000, 57, 1199–1205. [Google Scholar]
  258. Lopreato, G.F.; Phelan, R.; Borghese, C.M.; Beckstead, M.J.; Mihic, S.J. Inhaled Drugs of Abuse Enhance Serotonin-3 Receptor Function. Drug Alcohol. Depend. 2003, 70, 11–15. [Google Scholar] [CrossRef]
  259. Bale, A.S.; Meacham, C.A.; Benignus, V.A.; Bushnell, P.J.; Shafer, T.J. Volatile Organic Compounds Inhibit Human and Rat Neuronal Nicotinic Acetylcholine Receptors Expressed in Xenopus Oocytes. Toxicol. Appl. Pharmacol. 2005, 205, 77–88. [Google Scholar] [CrossRef]
  260. Venet, T.; Rumeau, C.; Campo, P.; Rieger, B.; Thomas, A.; Cour, C. Neuronal Circuits Involved in the Middle-Ear Acoustic Reflex. Toxicol. Sci. 2011, 119, 146–155. [Google Scholar] [CrossRef]
  261. Maguin, K.; Campo, P.; Parietti-Winkler, C. Toluene Can Perturb the Neuronal Voltage-Dependent Ca2+ Channels Involved in the Middle-Ear Reflex. Toxicol. Sci. 2009, 107, 473–481. [Google Scholar] [CrossRef]
  262. Besser, R.; Krämer, G.; Thümler, R.; Bohl, J.; Gutmann, L.; Hopf, H.C. Acute Trimethyltin Limbic-cerebellar Syndrome. Neurology 1987, 37, 945. [Google Scholar] [CrossRef]
  263. Yamasoba, T.; Goto, Y.; Komaki, H.; Mimaki, M.; Sudo, A.; Suzuki, M. Cochlear Damage Due to Germanium-Induced Mitochondrial Dysfunction in Guinea Pigs. Neurosci. Lett. 2006, 395, 18–22. [Google Scholar] [CrossRef]
  264. Crofton, K.M.; Ding, D.-L.; Padich, R.; Taylor, M.; Henderson, D. Hearing Loss Following Exposure during Development to Polychlorinated Biphenyls: A Cochlear Site of Action. Hear. Res. 2000, 144, 196–204. [Google Scholar] [CrossRef]
  265. Safe, T.M.; Luebke, A.E. Prenatal Low Dosage Dioxin (TCDD) Exposure Impairs Cochlear Function Resulting in Auditory Neuropathy. Hear. Res. 2016, 331, 7–12. [Google Scholar] [CrossRef]
  266. Woolley, D.E. Toxicological and Pharmacological Studies of Visual and Auditory Potentials Evoked in the Cerebellum of the Rat. Proc. West. Pharmacol. Soc. 1968, 11, 69–73. [Google Scholar]
  267. Hadjab, S.; Maurel, D.; Cazals, Y.; Siaud, P. Hexachlorobenzene, a Dioxin-like Compound, Disrupts Auditory Function in Rat. Hear. Res. 2004, 191, 125–134. [Google Scholar] [CrossRef]
  268. Meherali, S.; Campbell, A.; Hartling, L.; Scott, S. Understanding Parents’ Experiences and Information Needs on Pediatric Acute Otitis Media: A Qualitative Study. J. Patient Exp. 2019, 6, 53–61. [Google Scholar] [CrossRef]
  269. Rosenfeld, R.M.; Kay, D. Natural History of Untreated Otitis Media. Laryngoscope 2003, 113, 1645–1657. [Google Scholar] [CrossRef]
  270. Verhoeff, M.; Van Der Veen, E.L.; Rovers, M.M.; Sanders, E.A.M.; Schilder, A.G.M. Chronic Suppurative Otitis Media: A Review. Int. J. Pediatr. Otorhinolaryngol. 2006, 70, 1–12. [Google Scholar] [CrossRef] [PubMed]
  271. Yilmaz, T.; Koçcan, E.G.; Besler, H.T.; Yilmaz, G.; Gürsel, B. The Role of Oxidants and Antioxidants in Otitis Media with Effusion in Children. Otolaryngol.—Head Neck Surg. 2004, 131, 797–803. [Google Scholar] [CrossRef]
  272. Sagiroglu, S.; Ates, S.; Tolun, F.; Oztarakci, H. Evaluation of Oxidative Stress and Antioxidants Effect on Turning Process Acute Otitis Media to Chronic Otitis Media with Effusion. Niger. J. Clin. Pract. 2019, 22, 375. [Google Scholar] [CrossRef]
  273. Zhou, R.; Yazdi, A.S.; Menu, P.; Tschopp, J. A Role for Mitochondria in NLRP3 Inflammasome Activation. Nature 2011, 469, 221–225. [Google Scholar] [CrossRef] [PubMed]
  274. Zhang, Y.; Yang, W.; Li, W.; Zhao, Y. NLRP3 Inflammasome: Checkpoint Connecting Innate and Adaptive Immunity in Autoimmune Diseases. Front. Immunol. 2021, 12, 732933. [Google Scholar] [CrossRef]
  275. Zhong, Z.; Zhai, Y.; Liang, S.; Mori, Y.; Han, R.; Sutterwala, F.S.; Qiao, L. TRPM2 Links Oxidative Stress to NLRP3 Inflammasome Activation. Nat. Commun. 2013, 4, 1611. [Google Scholar] [CrossRef] [PubMed]
  276. Li, P.; Li, S.; Wang, L.; Li, H.; Wang, Y.; Liu, H.; Wang, X.; Zhu, X.; Liu, Z.; Ye, F.; et al. Mitochondrial Dysfunction in Hearing Loss: Oxidative Stress, Autophagy and NLRP3 Inflammasome. Front. Cell Dev. Biol. 2023, 11, 1119773. [Google Scholar] [CrossRef]
  277. Mizushima, Y.; Fujimoto, C.; Kashio, A.; Kondo, K.; Yamasoba, T. Macrophage Recruitment, but Not Interleukin 1 Beta Activation, Enhances Noise-Induced Hearing Damage. Biochem. Biophys. Res. Commun. 2017, 493, 894–900. [Google Scholar] [CrossRef] [PubMed]
  278. Paciello, F.; Fetoni, A.R.; Rolesi, R.; Wright, M.B.; Grassi, C.; Troiani, D.; Paludetti, G. Pioglitazone Represents an Effective Therapeutic Target in Preventing Oxidative/Inflammatory Cochlear Damage Induced by Noise Exposure. Front. Pharmacol. 2018, 9, 1103. [Google Scholar] [CrossRef]
  279. Wakabayashi, K.; Fujioka, M.; Kanzaki, S.; Okano, H.J.; Shibata, S.; Yamashita, D.; Masuda, M.; Mihara, M.; Ohsugi, Y.; Ogawa, K.; et al. Blockade of Interleukin-6 Signaling Suppressed Cochlear Inflammatory Response and Improved Hearing Impairment in Noise-Damaged Mice Cochlea. Neurosci. Res. 2010, 66, 345–352. [Google Scholar] [CrossRef] [PubMed]
  280. Giorgi, C.; Marchi, S.; Simoes, I.C.M.; Ren, Z.; Morciano, G.; Perrone, M.; Patalas-Krawczyk, P.; Borchard, S.; Jędrak, P.; Pierzynowska, K.; et al. Mitochondria and Reactive Oxygen Species in Aging and Age-Related Diseases. Int. Rev. Cell Mol. Biol. 2018, 340, 209–344. [Google Scholar] [CrossRef]
  281. Srivastava, S. The Mitochondrial Basis of Aging and Age-Related Disorders. Genes 2017, 8, 398. [Google Scholar] [CrossRef] [PubMed]
  282. Liguori, I.; Russo, G.; Curcio, F.; Bulli, G.; Aran, L.; Della-Morte, D.; Gargiulo, G.; Testa, G.; Cacciatore, F.; Bonaduce, D.; et al. Oxidative Stress, Aging, and Diseases. Clin. Interv. Aging 2018, 13, 757–772. [Google Scholar] [CrossRef]
  283. Trist, B.G.; Hilton, J.B.; Hare, D.J.; Crouch, P.J.; Double, K.L. Superoxide Dismutase 1 in Health and Disease: How a Frontline Antioxidant Becomes Neurotoxic. Angew. Chem. Int. Ed. 2021, 60, 9215–9246. [Google Scholar] [CrossRef]
  284. McFadden, S.L.; Ding, D.; Reaume, A.G.; Flood, D.G.; Salvi, R.J. Age-Related Cochlear Hair Cell Loss Is Enhanced in Mice Lacking Copper/Zinc Superoxide Dismutase. Neurobiol. Aging 1999, 20, 1–8. [Google Scholar] [CrossRef]
  285. Keithley, E.M.; Canto, C.; Zheng, Q.Y.; Wang, X.; Fischel-Ghodsian, N.; Johnson, K.R. Cu/Zn Superoxide Dismutase and Age-Related Hearing Loss. Hear. Res. 2005, 209, 76–85. [Google Scholar] [CrossRef]
  286. Ohlemiller, K.K.; McFadden, S.L.; Ding, D.-L.; Lear, P.M.; Ho, Y.-S. Targeted Mutation of the Gene for Cellular Glutathione Peroxidase (Gpx1) Increases Noise-Induced Hearing Loss in Mice. JARO 2000, 1, 243–254. [Google Scholar] [CrossRef]
  287. Nissanka, N.; Moraes, C.T. Mitochondrial DNA Damage and Reactive Oxygen Species in Neurodegenerative Disease. FEBS Lett. 2018, 592, 728–742. [Google Scholar] [CrossRef]
  288. Yang, X.; Zhang, R.; Nakahira, K.; Gu, Z. Mitochondrial DNA Mutation, Diseases, and Nutrient-Regulated Mitophagy. Annu. Rev. Nutr. 2019, 39, 201–226. [Google Scholar] [CrossRef] [PubMed]
  289. Hahn, A.; Zuryn, S. Mitochondrial Genome (mtDNA) Mutations That Generate Reactive Oxygen Species. Antioxidants 2019, 8, 392. [Google Scholar] [CrossRef] [PubMed]
  290. Kujoth, G.C.; Hiona, A.; Pugh, T.D.; Someya, S.; Panzer, K.; Wohlgemuth, S.E.; Hofer, T.; Seo, A.Y.; Sullivan, R.; Jobling, W.A.; et al. Mitochondrial DNA Mutations, Oxidative Stress, and Apoptosis in Mammalian Aging. Science 2005, 309, 481–484. [Google Scholar] [CrossRef] [PubMed]
  291. Guo, D.; Zhang, A.; Zou, T.; Ding, R.; Chen, K.; Pan, Y.; Ji, P.; Ye, B.; Xiang, M. The Influence of Metabolic Syndrome on Age-Related Hearing Loss from the Perspective of Mitochondrial Dysfunction. Front. Aging Neurosci. 2022, 14, 930105. [Google Scholar] [CrossRef] [PubMed]
  292. Zhang, L.; Wang, J.; Sun, H.; Feng, G.; Gao, Z. Interactions between the Hippocampus and the Auditory Pathway. Neurobiol. Learn. Mem. 2022, 189, 107589. [Google Scholar] [CrossRef]
  293. Chen, A.P.F.; Malgady, J.M.; Chen, L.; Shi, K.W.; Cheng, E.; Plotkin, J.L.; Ge, S.; Xiong, Q. Nigrostriatal Dopamine Pathway Regulates Auditory Discrimination Behavior. Nat. Commun. 2022, 13, 5942. [Google Scholar] [CrossRef]
  294. Bonna, K.; Finc, K.; Zimmermann, M.; Bola, L.; Mostowski, P.; Szul, M.; Rutkowski, P.; Duch, W.; Marchewka, A.; Jednoróg, K.; et al. Early Deafness Leads to Re-Shaping of Functional Connectivity beyond the Auditory Cortex. Brain Imaging Behav. 2021, 15, 1469–1482. [Google Scholar] [CrossRef] [PubMed]
  295. Wang, X.; Xu, P.; Li, P.; Wang, Z.; Zhao, F.; Gao, Z.; Xu, L.; Luo, Y.; Fan, J.; Liu, P. Alterations in Gray Matter Volume Due to Unilateral Hearing Loss. Sci. Rep. 2016, 6, 25811. [Google Scholar] [CrossRef]
  296. Jafari, Z.; Kolb, B.E.; Mohajerani, M.H. Auditory Dysfunction in Parkinson’s Disease. Mov. Disord. 2020, 35, 537–550. [Google Scholar] [CrossRef]
  297. Eversfield, C.L.; Orton, L.D. Auditory and Visual Hallucination Prevalence in Parkinson’s Disease and Dementia with Lewy Bodies: A Systematic Review and Meta-Analysis. Psychol. Med. 2019, 49, 2342–2353. [Google Scholar] [CrossRef] [PubMed]
  298. Golden, H.L.; Nicholas, J.M.; Yong, K.X.X.; Downey, L.E.; Schott, J.M.; Mummery, C.J.; Crutch, S.J.; Warren, J.D. Auditory Spatial Processing in Alzheimer’s Disease. Brain 2015, 138, 189–202. [Google Scholar] [CrossRef]
  299. Anshu, P.; Jadhav, D.; Agrawal, S.; Durge, V.; Jain, N.; Ravat, S. Case Report of SOD1 ALS Presenting with Prominent Hearing Impairment. J. Neurol. Sci. 2021, 429, 118345. [Google Scholar] [CrossRef]
  300. Wang, H.-F.; Zhang, W.; Rolls, E.T.; Li, Y.; Wang, L.; Ma, Y.-H.; Kang, J.; Feng, J.; Yu, J.-T.; Cheng, W. Hearing Impairment Is Associated with Cognitive Decline, Brain Atrophy and Tau Pathology. eBioMedicine 2022, 86, 104336. [Google Scholar] [CrossRef]
  301. Rutherford, B.R.; Brewster, K.; Golub, J.S.; Kim, A.H.; Roose, S.P. Sensation and Psychiatry: Linking Age-Related Hearing Loss to Late-Life Depression and Cognitive Decline. AJP 2018, 175, 215–224. [Google Scholar] [CrossRef] [PubMed]
  302. Linszen, M.M.J.; Brouwer, R.M.; Heringa, S.M.; Sommer, I.E. Increased Risk of Psychosis in Patients with Hearing Impairment: Review and Meta-Analyses. Neurosci. Biobehav. Rev. 2016, 62, 1–20. [Google Scholar] [CrossRef] [PubMed]
  303. Kadiiska, M.B.; Basu, S.; Brot, N.; Cooper, C.; Saari Csallany, A.; Davies, M.J.; George, M.M.; Murray, D.M.; Jackson Roberts, L.; Shigenaga, M.K.; et al. Biomarkers of Oxidative Stress Study V: Ozone Exposure of Rats and Its Effect on Lipids, Proteins, and DNA in Plasma and Urine. Free Radic. Biol. Med. 2013, 61, 408–415. [Google Scholar] [CrossRef] [PubMed]
  304. Pak, J.H.; Kim, Y.; Yi, J.; Chung, J.W. Antioxidant Therapy against Oxidative Damage of the Inner Ear: Protection and Preconditioning. Antioxidants 2020, 9, 1076. [Google Scholar] [CrossRef]
  305. Hazlitt, R.A.; Min, J.; Zuo, J. Progress in the Development of Preventative Drugs for Cisplatin-Induced Hearing Loss: Miniperspective. J. Med. Chem. 2018, 61, 5512–5524. [Google Scholar] [CrossRef]
  306. Molina, S.J.; Miceli, M.; Guelman, L.R. Noise Exposure and Oxidative Balance in Auditory and Extra-Auditory Structures in Adult and Developing Animals. Pharmacological Approaches Aimed to Minimize Its Effects. Pharmacol. Res. 2016, 109, 86–91. [Google Scholar] [CrossRef]
  307. Kopke, R.; Slade, M.D.; Jackson, R.; Hammill, T.; Fausti, S.; Lonsbury-Martin, B.; Sanderson, A.; Dreisbach, L.; Rabinowitz, P.; Torre, P.; et al. Efficacy and Safety of N-Acetylcysteine in Prevention of Noise Induced Hearing Loss: A Randomized Clinical Trial. Hear. Res. 2015, 323, 40–50. [Google Scholar] [CrossRef] [PubMed]
  308. Kang, W.-S.; Chung, J.-W. Ingestion of Korean Red Ginseng after Noise Exposure Can Potentiate Rapid Recovery of Hearing in Mice. J. Ginseng Res. 2010, 34, 336–341. [Google Scholar] [CrossRef]
  309. Cunningham, L.L.; Tucci, D.L. Hearing Loss in Adults. N. Engl. J. Med. 2017, 377, 2465–2473. [Google Scholar] [CrossRef] [PubMed]
  310. MacDonald, R.S. The Role of Zinc in Growth and Cell Proliferation. J. Nutr. 2000, 130, 1500S–1508S. [Google Scholar] [CrossRef] [PubMed]
  311. Yi, J.; Chung, J.W.; Pak, J.H. Zinc Is an Essential Element for the Maintenance of Redox Homeostasis and Cell Cycle in Murine Auditory Hair Cells. J. Nutr. Biochem. 2022, 100, 108901. [Google Scholar] [CrossRef]
  312. Kamogashira, T.; Fujimoto, C.; Yamasoba, T. Reactive Oxygen Species, Apoptosis, and Mitochondrial Dysfunction in Hearing Loss. BioMed Res. Int. 2015, 2015, 1–7. [Google Scholar] [CrossRef] [PubMed]
  313. Lopez-Novoa, J.M.; Quiros, Y.; Vicente, L.; Morales, A.I.; Lopez-Hernandez, F.J. New Insights into the Mechanism of Aminoglycoside Nephrotoxicity: An Integrative Point of View. Kidney Int. 2011, 79, 33–45. [Google Scholar] [CrossRef]
  314. Vokes, E.E. Induction Chemotherapy for Head and Neck Cancer: Recent Data. Oncologist 2010, 15, 3–7. [Google Scholar] [CrossRef] [PubMed]
  315. Alexander, T.H.; Harris, J.P. Incidence of Sudden Sensorineural Hearing Loss. Otol. Neurotol. 2013, 34, 1586–1589. [Google Scholar] [CrossRef]
  316. Mukherjea, D.; Jajoo, S.; Sheehan, K.; Kaur, T.; Sheth, S.; Bunch, J.; Perro, C.; Rybak, L.P.; Ramkumar, V. NOX3 NADPH Oxidase Couples Transient Receptor Potential Vanilloid 1 to Signal Transducer and Activator of Transcription 1-Mediated Inflammation and Hearing Loss. Antioxid. Redox Signal. 2011, 14, 999–1010. [Google Scholar] [CrossRef]
  317. Baek, J.-I.; Kim, Y.-R.; Lee, K.-Y.; Kim, U.-K. Mitochondrial Redox System: A Key Target of Antioxidant Therapy to Prevent Acquired Sensorineural Hearing Loss. Front. Pharmacol. 2023, 14, 1176881. [Google Scholar] [CrossRef] [PubMed]
  318. Freyer, D.R.; Chen, L.; Krailo, M.D.; Knight, K.; Villaluna, D.; Bliss, B.; Pollock, B.H.; Ramdas, J.; Lange, B.; Van Hoff, D.; et al. Effects of Sodium Thiosulfate versus Observation on Development of Cisplatin-Induced Hearing Loss in Children with Cancer (ACCL0431): A Multicentre, Randomised, Controlled, Open-Label, Phase 3 Trial. Lancet Oncol. 2017, 18, 63–74. [Google Scholar] [CrossRef] [PubMed]
  319. Brock, P.R.; Maibach, R.; Childs, M.; Rajput, K.; Roebuck, D.; Sullivan, M.J.; Laithier, V.; Ronghe, M.; Dall’Igna, P.; Hiyama, E.; et al. Sodium Thiosulfate for Protection from Cisplatin-Induced Hearing Loss. N. Engl. J. Med. 2018, 378, 2376–2385. [Google Scholar] [CrossRef] [PubMed]
  320. Dickey, D.T.; Wu, Y.J.; Muldoon, L.L.; Neuwelt, E.A. Protection against Cisplatin-Induced Toxicities by N-Acetylcysteine and Sodium Thiosulfate as Assessed at the Molecular, Cellular, and in Vivo Levels. J. Pharmacol. Exp. Ther. 2005, 314, 1052–1058. [Google Scholar] [CrossRef] [PubMed]
  321. Hu, S.; Sun, Q.; Xu, F.; Jiang, N.; Gao, J. Age-Related Hearing Loss and Its Potential Drug Candidates: A Systematic Review. Chin. Med. 2023, 18, 121. [Google Scholar] [CrossRef] [PubMed]
  322. Pisani, A.; Paciello, F.; Montuoro, R.; Rolesi, R.; Galli, J.; Fetoni, A.R. Antioxidant Therapy as an Effective Strategy against Noise-Induced Hearing Loss: From Experimental Models to Clinic. Life 2023, 13, 1035. [Google Scholar] [CrossRef]
  323. Bast, A.; Haenen, G.R.M.M. The Toxicity of Antioxidants and Their Metabolites. Environ. Toxicol. Pharmacol. 2002, 11, 251–258. [Google Scholar] [CrossRef]
Figure 1. The anatomical and biological structure of the auditory system. (A) The anatomy of the auditory system comprises the outer auditory canal (outer ear), auditory ossicles (middle ear), and the cochlea (inner ear). (B) The outer ear canal forms a physical passageway which collects and channels external sound signals into the middle ear. (C) The sound waves are made up of high-amplitude vibrations of air particles (represented by a black waveform with an arrow showing the direction of movement) which displace the eardrum from its basal position (top compared with bottom figure, red box). The physical connection between the eardrum, ossicles, and tympanic membrane allows physical conduction of the movement of the eardrum to the tympanic membrane and the inner ear of the auditory system. (D) The inner ear is made up of the fluid-filled, spiral-shaped cochlea, which contains the mechanosensory hair cells required for the transduction of auditory signals into the CNS (top left). The stereocilia on hair cells are connected to the tectorial membrane through the calcium-rich filamentous structures that allow the transmission of fluid-dependent movement of the tectorial membrane to the mechanosensitive mechanoelectrical transduction (MET) channel of the hair cells. Activation of the sensory hair cells leads to the release of neurotransmitters targeting the innervating auditory nerve for transmission of auditory signals into the CNS (top right). The inner and outer hair cells comprise slightly different molecular components, which allow their specialized functions (bottom).
Figure 1. The anatomical and biological structure of the auditory system. (A) The anatomy of the auditory system comprises the outer auditory canal (outer ear), auditory ossicles (middle ear), and the cochlea (inner ear). (B) The outer ear canal forms a physical passageway which collects and channels external sound signals into the middle ear. (C) The sound waves are made up of high-amplitude vibrations of air particles (represented by a black waveform with an arrow showing the direction of movement) which displace the eardrum from its basal position (top compared with bottom figure, red box). The physical connection between the eardrum, ossicles, and tympanic membrane allows physical conduction of the movement of the eardrum to the tympanic membrane and the inner ear of the auditory system. (D) The inner ear is made up of the fluid-filled, spiral-shaped cochlea, which contains the mechanosensory hair cells required for the transduction of auditory signals into the CNS (top left). The stereocilia on hair cells are connected to the tectorial membrane through the calcium-rich filamentous structures that allow the transmission of fluid-dependent movement of the tectorial membrane to the mechanosensitive mechanoelectrical transduction (MET) channel of the hair cells. Activation of the sensory hair cells leads to the release of neurotransmitters targeting the innervating auditory nerve for transmission of auditory signals into the CNS (top right). The inner and outer hair cells comprise slightly different molecular components, which allow their specialized functions (bottom).
Antioxidants 13 00598 g001
Figure 2. The structural and functional impact of hereditary mutations in genes involved in auditory function. Myosin VI (Myo6) and alpha-tectorin (TECTA) are involved in stereocilia structure maintenance. Mutations in Myo6 and TECTA lead to the mislocalization of the actin (MYO, bottom right) and the loss in alignment of the stereocilia structure to the tectorial membrane (TECTA, bottom right). Alternatively, mutations in the functional components of the hair cells can lead to significant alteration of the ionic gradient required for maintenance of the endolymphatic potential (GJB, right) and hair cell membrane potential (KCNQ4, right), which impacts the ease of hair cell activation and signal transmission to the auditory nerve. OTOF and SLC17A8 mutations are linked to a reduction in the quantal size of the neurotransmitter released (SLC17A8, right) and loss of the abilities of vesicle fusion and neurotransmitter release (OTOF, bottom right) upon IHC activation. Changes to the OHC electromobility (SLC26A5, depolarization and hyperpolarization) further reduce the sensitivity of the sensory hair cells toward stimulus-evoked activation. The contributions of individual effects or the combinatorial effect of the altered structural and functional components reduce the sensitivity and, in severe cases, ablate stimulus-dependent hair cell activation, leading to hearing impairment.
Figure 2. The structural and functional impact of hereditary mutations in genes involved in auditory function. Myosin VI (Myo6) and alpha-tectorin (TECTA) are involved in stereocilia structure maintenance. Mutations in Myo6 and TECTA lead to the mislocalization of the actin (MYO, bottom right) and the loss in alignment of the stereocilia structure to the tectorial membrane (TECTA, bottom right). Alternatively, mutations in the functional components of the hair cells can lead to significant alteration of the ionic gradient required for maintenance of the endolymphatic potential (GJB, right) and hair cell membrane potential (KCNQ4, right), which impacts the ease of hair cell activation and signal transmission to the auditory nerve. OTOF and SLC17A8 mutations are linked to a reduction in the quantal size of the neurotransmitter released (SLC17A8, right) and loss of the abilities of vesicle fusion and neurotransmitter release (OTOF, bottom right) upon IHC activation. Changes to the OHC electromobility (SLC26A5, depolarization and hyperpolarization) further reduce the sensitivity of the sensory hair cells toward stimulus-evoked activation. The contributions of individual effects or the combinatorial effect of the altered structural and functional components reduce the sensitivity and, in severe cases, ablate stimulus-dependent hair cell activation, leading to hearing impairment.
Antioxidants 13 00598 g002
Figure 3. Progressive loss of cells and auditory function with time. Genetic predisposition to auditory disfunction and cumulative exposure toward ototoxic stimulus leads to the gradual loss of sensory hair cells over time.
Figure 3. Progressive loss of cells and auditory function with time. Genetic predisposition to auditory disfunction and cumulative exposure toward ototoxic stimulus leads to the gradual loss of sensory hair cells over time.
Antioxidants 13 00598 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Yeo, X.Y.; Kwon, S.; Rinai, K.R.; Lee, S.; Jung, S.; Park, R. A Consolidated Understanding of the Contribution of Redox Dysregulation in the Development of Hearing Impairment. Antioxidants 2024, 13, 598. https://doi.org/10.3390/antiox13050598

AMA Style

Yeo XY, Kwon S, Rinai KR, Lee S, Jung S, Park R. A Consolidated Understanding of the Contribution of Redox Dysregulation in the Development of Hearing Impairment. Antioxidants. 2024; 13(5):598. https://doi.org/10.3390/antiox13050598

Chicago/Turabian Style

Yeo, Xin Yi, Soohyun Kwon, Kimberley R. Rinai, Sungsu Lee, Sangyong Jung, and Raekil Park. 2024. "A Consolidated Understanding of the Contribution of Redox Dysregulation in the Development of Hearing Impairment" Antioxidants 13, no. 5: 598. https://doi.org/10.3390/antiox13050598

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