Next Article in Journal
Toward a Platform for the Treatment of Burns: An Assessment of Nanoemulsions vs. Nanostructured Lipid Carriers Loaded with Curcumin
Previous Article in Journal
Expression of Serpin Family E Member 1 (SERPINE1) Is Associated with Poor Prognosis of Gastric Adenocarcinoma
Previous Article in Special Issue
Novel Pathogenic Variants in the Gene Encoding Stereocilin (STRC) Causing Non-Syndromic Moderate Hearing Loss in Spanish and Argentinean Subjects
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Therapeutic Progress and Future Perspectives for the Treatment of Hearing Loss

by
Joey Lye
1,2,†,
Derek S. Delaney
1,3,†,
Fiona K. Leith
1,2,
Varda S. Sardesai
1,
Samuel McLenachan
4,5,
Fred K. Chen
4,5,6,7,8,
Marcus D. Atlas
1,2 and
Elaine Y. M. Wong
1,2,9,*
1
Hearing Therapeutics, Ear Science Institute Australia, Nedlands, WA 6009, Australia
2
Centre for Ear Sciences, Medical School, The University of Western Australia, Nedlands, WA 6009, Australia
3
Faculty of Health Sciences, Curtin Health Innovation Research Institute, Curtin University, Bentley, WA 6102, Australia
4
Ocular Tissue Engineering Laboratory, Lions Eye Institute, Nedlands, WA 6009, Australia
5
Centre for Ophthalmology and Visual Sciences, The University of Western Australia, Nedlands, WA 6009, Australia
6
Vitroretinal Surgery, Royal Perth Hospital, Perth, WA 6000, Australia
7
Ophthalmology, Department of Surgery, University of Melbourne, East Melbourne, VIC 3002, Australia
8
Centre for Eye Research Australia, Royal Victorian Eye and Ear Hospital, East Melbourne, VIC 3002, Australia
9
Curtin Medical School, Faculty of Health Sciences, Curtin University, Bentley, WA 6102, Australia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Biomedicines 2023, 11(12), 3347; https://doi.org/10.3390/biomedicines11123347
Submission received: 16 November 2023 / Revised: 6 December 2023 / Accepted: 11 December 2023 / Published: 18 December 2023
(This article belongs to the Special Issue Genetic Research on Hearing Loss 2.0)

Abstract

:
Up to 1.5 billion people worldwide suffer from various forms of hearing loss, with an additional 1.1 billion people at risk from various insults such as increased consumption of recreational noise-emitting devices and ageing. The most common type of hearing impairment is sensorineural hearing loss caused by the degeneration or malfunction of cochlear hair cells or spiral ganglion nerves in the inner ear. There is currently no cure for hearing loss. However, emerging frontier technologies such as gene, drug or cell-based therapies offer hope for an effective cure. In this review, we discuss the current therapeutic progress for the treatment of hearing loss. We describe and evaluate the major therapeutic approaches being applied to hearing loss and summarize the key trials and studies.

1. Introduction

1.1. Clinical Burden of Sensorineural Hearing Loss

The World Health Organization (WHO) has estimated over 1.5 billion people worldwide currently experience some form of hearing loss (HL), which is expected to increase significantly by 2050 [1]. Hearing impairment is a disabling disorder which can significantly affect livelihood and is a leading global public health problem. The economic burden of HL has been estimated to cost $980 billion USD annually to provide healthcare, societal and educational support to patients [1]. A study reported by Mohr et al. [2] suggested that an individual living in the United States is expected to spend $297,000 USD over their lifetime on hearing health. However, a recent study suggested the cost of hearing-related health expenditures may exceed $1 million USD for at-need children to have access to special education, counselling and hearing interventions [3].
People with HL can experience significant psychosocial impacts throughout their lives. When HL is unaddressed in early life, educational outcomes and the development of social skills are impacted, which can lead to reduced employment opportunities and quality of life in adulthood [4,5]. Most cases of HL are preventable but limited awareness and delayed self-diagnosis contribute to the neglect of early intervention and preventative measures. Furthermore, the prevalence of HL is known to increase with age, with age-related hearing loss (ARHL) or presbycusis being a common problem for older people. In recent years, there have been increased reports of a significant association of cognitive impairment and ARHL [6,7]. Dementia has been shown to be up to five times more prevalent among older people with moderate to severe HL than those with normal hearing [8].
Sensorineural hearing loss (SNHL) is the most frequent form of HL and is defined by the inability to transmit incoming soundwaves and perceive them as sound due to pathophysiological changes in the auditory pathway. This is defined by the degeneration of cochlear hair cells (HCs) and/or afferent neurons and sounds are perceived as muted and distorted compared to other forms of HL. SNHL is progressive and results in permanent and irreparable damage. SNHL can be inherited or acquired through age, genetics and environmental factors, which can also interact to cause unilateral or bilateral HL [9].

1.2. Cause of Sensorineural Hearing Loss

The cochlea is responsible for the transduction of sound to the brain and the sensory cells of the organ of Corti within it facilitate this process. The organ of Corti houses inner hair cells (IHCs) and outer hair cells (OHCs), which are responsible for the primary transduction of sound through spiral ganglion nerves (SGNs) and mediating oscillatory forces, respectively [10] (Figure 1). HCs are named for the “hair bundle” on their apical surface, which is comprised of actin-based projections called stereocilia. Stereocilia are embedded in the tectorial membrane, through which they detect soundwaves originating from the middle ear [11]. Soundwaves cause the tectorial membrane to vibrate, which moves stereocilia and triggers the opening of mechanoelectrical transduction channels, causing depolarization of the SGNs and subsequently, transmission of the sound signal to the brain [11,12,13,14,15]. Moreover, HCs are surrounded by a heterogenous population of various types of supporting cells (SCs), which are thought to play a role in potassium ion regulation, osmotic balance and limited epithelial repair following trauma (Figure 1).
SNHL is a result of damage to or malfunction of the organ of Corti that interferes with the process of sound transduction. SNHL can be acquired or hereditary. Acquired SNHL is largely environmental and is a result of cumulative damage to the organ of Corti over the course of life [16]. Noise exposure from industrial work or loud music are common causes of noise-induced hearing loss (NIHL). Drug-induced HL is a significant clinical problem as ototoxicity is a prevalent side effect, with a review by Rizk and colleagues in 2020 listing 194 medications of various classes having ototoxic side effects [17]. Some of these medications that are prescribed for long-term management of chronic disorders, such as gentamicin or cisplatin, can cause severe and irreversible damage to sensory cells in the cochlea.
ARHL, or presbycusis, and NIHL will become increasingly prevalent with the extending human lifespan and increased use of personal listening devices [1]. Bilateral SNHL currently affects over 65% of adults over 65 years of age [16]. Additionally, smoking and cardiovascular and metabolic disorders have been shown to be associated with SNHL. The causal mechanisms are still to be established for these associations, however it is thought for cardiovascular and metabolic disorders that damage to the cochlear vasculature may play a role [18,19,20,21].
Hereditary SNHL affects around 1 in every 1000 newborns and can be devastating with genetic mutations to single genes, resulting in complete or progressive deafness in some diseases [22,23]. Many of the genes affected in disorders of hereditary SNHL encode the essential components of HCs and mechanoelectrical transduction. For example, GJB2 encodes the gap junction protein connexin 26, while many of the genes in Usher syndrome such as MYO7A and WHRN encode for essential hair bundle components [24,25]. Mutations in these genes severely affect signal transduction, as unbalanced ion concentrations and destabilized stereocilia render HCs insensitive, or unresponsive in worst cases, to sound and result in deafness. There are currently 124 genes that have been implicated in various forms of non-syndromic hereditary SNHL (https://hereditaryhearingloss.org/).

1.3. Current Interventions against Sensorineural Hearing Loss

The most common intervention for patients with SNHL is using prosthetics such as hearing aids or cochlear implants to aid hearing. While these are effective for less severe forms of SNHL, they do still require some HC functionality for sound transduction [26]. Moreover, they are currently unable to mimic the quality of natural hearing.
There is a severe paucity of small molecule drugs available for treating SNHL and these are known to have varying degrees of effectiveness. The corticosteroid dexamethasone is commonly administered either orally or intratympanically for its anti-inflammatory properties and has shown some benefit [27,28]. Unfortunately, dexamethasone has poor bioavailability in the cochlea due to limited uptake in the inner ear when administered orally and is readily cleared from the cochlear fluids when injected intratympanically [29]. Recently, SPT-2101, a hydrogel formulation containing dexamethasone, has been developed by Spiral Therapeutics (South Francisco, CA, USA) and is currently undergoing a Phase I clinical trial in Australia (ACTRN12621000964819), aiming to provide sustained release in the inner ear. For drug-induced HL, sodium thiosulfate was recently approved by the US Food and Drug Administration (FDA) to prevent cisplatin-induced HL in patients undergoing cisplatin chemotherapy [30]. Sodium thiosulfate directly chelates cisplatin, however, which risks interfering with the chemotherapeutic effects [31]. A clinical trial of intratympanically administered sodium thiosulfate, using a hydrogel-based formulation for controlled release (DB-020, Decibel Therapeutics, Boston, MA, USA), recently completed Phase I clinical trials (Table 1). Intravenously administered sodium thiosulfate remains the only novel drug to be approved by the US FDA in 20 years of research in hearing therapeutics [32].

2. Novel Therapeutic Approaches to Restore Hearing

While the current options of drug treatment for any form of SNHL are limited, there are many novel treatments undergoing clinical trials. As corticosteroid therapy is proving to be insufficient for widely treating SNHL, novel drugs targeting other pathways are being investigated. One pathway which has recently seen great research interest and investment is regenerating HCs through transdifferentiation of SCs. Small molecule and genetic approaches are being developed to induce the signaling pathways to initiate HC transdifferentiation. Moreover, genetic approaches are being increasingly investigated to treat the most severe hereditary forms of SNHL where, for example, gene supplementation or editing of an endogenously faulty gene could restore HC function. In the following sections, we will review the recent therapeutic strategies for targeting the pathways that promote the restoration of the sensory epithelium and SGNs.

3. Emerging Drug Therapies

As with other highly specialized tissues, the development of the mammalian inner ear is guided by the precise spatiotemporal activation and inhibition of key signaling pathways. These key signaling pathways include Wnt, which interacts with Sonic Hedgehog signaling to specify the apical–basal polarity of the cochlea and cochlear outgrowth, and Notch, which is involved in cell fate specification and patterning within the organ of Corti [33]. Several groups have demonstrated that in vitro manipulation of these signaling pathways can induce differentiation of embryonic and induced pluripotent stem cells into HC-like cells [34,35,36,37].
Most recently, Zheng-Yi Chen’s group was able to demonstrate a novel approach to reprogram SCs into HCs in adult mice using small molecule drugs. Quan et al. [37] demonstrated the transdifferentiation of primarily interdental cells into OHCs in adult mouse cochlear explants using a cocktail treatment consisting of valproic acid, siRNAs activating c-Myc, lithium chloride as a Wnt activator and a cyclic AMP agonist Forsklin. Single-cell RNAseq of the transformed HC-like cells showed that the cocktail stimulated a direct transition from an interdental cell type to that resembling fetal HCs, showed by the co-staining of the HC markers with SOX2 [37]. Moreover, this study achieved an additional major milestone by demonstrating transdifferentiation in the adult mouse cochlea, whereas most studies to date have been executed on neonatal mice.
A few small molecule drugs attempting to induce HC differentiation have recently undergone clinical trials for treating various forms of SNHL in humans. One is FX-322 from Frequency Therapeutics, which is an intratympanically injected formulation of the Wnt activator, CHIR99021, and histone deacetylase inhibitor, valproic acid. FX-322 just completed Phase IIb clinical trials, showing no statistically significant difference at day 90 compared to the placebo. The other drugs are LY3056480 and PIPE-505, from Audion Therapeutics (Amsterdam, The Netherlands) and Pipeline Therapeutics (San Diego, CA, USA), respectively, which are both γ-secretase inhibitors. γ-secretase plays a key role in Notch signaling by cleaving the intracellular domain of Notch and initiating the signaling pathway. Moreover, in the adult mammalian cochlea, Notch-mediated lateral inhibition prevents transdifferentiation of HCs [38,39]. At this stage, both studies appear to be either on hold or withdrawn.
Cellular quiescence occurs via the inhibition of cyclin/cyclin-dependent kinase (CDK) and the hypophosphorylation of retinoblastoma protein [40]. CDK inhibitors, such as p27Kip1, are expressed in the cochlea at a level sufficient enough to promote quiescence of the HCs [41,42,43]. The deletion of p27Kip1 in transgenic mice resulted in the proliferation of SCs and the regeneration of HCs [44,45,46]. The HC-specific conditional deletion of p27Kip1 in neonatal mice resulted in the proliferation and improved survival of HCs without any adverse effects on hearing, as measured by the auditory brainstem response (ABR). IHCs in particular were more proliferative than OHCs in response to the knockout of p27Kip1 [43]. These findings suggested that p27Kip1 could represent a good therapeutic target to minimize cell death and promote HC regeneration in the cochlea. Indeed, Sound Pharmaceuticals (Seattle, WA, USA) is conducting a preclinical study of SPI-5557, a drug that inhibits p27Kip1, with the aim of regenerating cochlear HCs [47].
Multiple strategies have been suggested to deliver pro-neurogenic growth factors to the inner ear to promote SGN growth and regeneration. Neurotrophins, such as the brain-derived neurotrophic factor or neurotrophin-3, have been demonstrated to have protective effects in animals exposed to ototoxic insults [48]. Moreover, they have also been shown to promote neural outgrowth and the formation of synapses, making them and their Trk receptors promising candidates for the treatment of SNHL [49,50]. Small molecule Trk receptor agonists have been proposed as a strategy to induce neuroprotective pathways and improve the pharmacokinetics of Trk receptor activation in the cochlea [49,51,52]. A recent study by Fernandez et al. [53] demonstrated the protective and regenerative effects of the TrkB agonists amitriptyline and 7,8-dihydroxyflavone on an animal model of glutamate excitotoxicity, which damages SGN terminal processes. Another recent study demonstrated that a neurotrophin-3 mimetic, 1Aa conjugated with risedronate, promoted the regeneration of cochlear ribbon synapses and neurite outgrowth in mouse cochlear explants [54]. An alternative approach for introducing neurotrophins into the cochlea has additionally been proposed through gene delivery to the cochlea [55]. We will more broadly discuss gene therapy in the following section.
While small molecule drugs are more traditional, stable and less complex than the newer therapies discussed in the subsequent sections, a major issue for therapy lies in drug delivery. The inner ear is among the hardest areas in the body to deliver drugs to, as systemically administered drugs are poorly taken up and the cochlea itself lies behind the tympanic and RWM when administering via the ear canal. Recent developments in hydrogel and nanotechnology have proposed to provide a reservoir of continually delivered drug and increase permeation. This would theoretically solve the problem of controlled release and continual dosage [56,57,58,59].
Although this review largely focuses on HC regeneration, there are many other small molecule inhibitors undergoing clinical trials. These are aimed at a range of SNHL etiologies, including sudden SNHL and cytomegalovirus-induced SNHL. While this review is more focused on applications for the more prevalent forms of acquired and hereditary SNHL, we have compiled a table of all the small molecule drugs currently undergoing clinical trials that is registered in the National Library of Medicine’s (NLM) database (www.clinicaltrials.gov) for other forms of SNHL, as shown in Table 1.

4. Gene Therapy

Gene therapy generally refers to the supplementation of an endogenously healthy gene and offers a prospective solution for certain forms of SNHL. In a landmark study, ATOH1 (also known as MATH1 for “mouse Atoh1” and HATH1 for “human ATOH1”), a basic helix-loop-helix transcription factor, was identified to be essential in the generation of inner ear HCs and was able to enhance HC regeneration when overexpressed [60,61]. ATOH1 has since been used as a prime candidate in HC regeneration research to varying degrees of success in published studies. Guinea pigs treated with ATOH1 gene therapy showed a significant increase in the number of HCs marked by a positive myosin VIIa (MYO7A) expression but failed to regenerate a full complement of ribbon synapses [62]. Overall, the hearing in these guinea pigs did not recover, suggesting that ATOH1 gene therapy alone is unable to convert non-sensory cells into HCs, and is not capable of rescuing the phenotype of surviving HCs [62]. Apart from ATOH1, other studies have looked at the candidate signaling pathways that control the ATOH1 gene expression, such as Notch signaling, which plays a key role in HC and SC differentiation, and Wnt/β-catenin signaling pathways.
A popular option for delivering gene therapy is using viral vectors, such as lentiviruses and adeno-associated viruses (AAVs). While lentiviral vectors have a greater packaging capacity than AAV vectors, ~7.5 kb versus ~5 kb, respectively, the genes that encode HC mechanotransduction components, commonly mutated or lost in genetic disorders such as Usher syndrome (USH), often exceed even the lentiviral packaging capacity.
The AAV delivery of full-length MYO7A cDNA has been shown to be effective in vivo and in vitro and resulted in wild-type levels of expression [63,64,65,66]. Packaging full-length MYO7A cDNA into AAVs produced vectors with heterogeneous, fragmented genomes (fAAVs) that were able to reconstitute full-length MYO7A cDNA post-infection. However, fAAVs are relatively ineffective and unable to be definitively characterized, so are unacceptable for clinical translation [65]. As MYO7A is relatively large, it exceeds the 5 kb capacity of a single AAV, so a dual AAV vector platform might overcome this limitation [64,65,66]. An alternative strategy to overcome the size limitations reported for OTOF (otoferlin) and PCDH15 (protocadherin 15) has been to decrease the gene size by deleting redundant domains, which has shown some success in mouse models [67,68].
Dual AAV vectors are designed as pairs, containing either the 3′ or 5′ fragments of MYO7A cDNA with an overlapping region. These AAVs can then be reconstituted by either splicing (trans-splicing), homologous recombination (overlapping) or a combination of the two methods (hybrid) [66]. Dual AAV treatments have transduced retinal pigment epithelium and photoreceptor cells, resulting in the expression of a functional full-length myosin VIIA protein and the rescue of the mutant phenotypes, which is suggestive of the successful homologous recombination of MYO7A gene fragments. However, this method had low efficiency and minimal overexpression of MYO7A in infected cells [64].
A breakthrough in rescuing HL and vestibular function has been achieved in adult mouse models via a synthetic designer AAV vector (AAV2/Anc80L65) [69]. The injection of AAVs into the posterior semicircular canal of adult mice has demonstrated efficient transduction of all IHCs, a proportion of OHCs, as well as all the SCs in the vestibular system. These findings indicated the potential of this viral vector and canalostomy as a surgical approach for gene therapy in cochlear and vestibular sensory organs [69].
In another study, AAV2/Anc80L65 injected in neonatal Ush1c c.216G>A mice revealed an improvement in IHC and OHC survival, with the HCs remaining mechanosensitive during the first postnatal week [70]. Using a similar approach, the injection of AAV2/1 vectors in Tmc1 mutant mice, known as Beethoven (Bth), targeted 80–90% of IHCs and yielded moderate auditory rescue [71].
The direct injection of AAV serotype 2/8 to deliver DFNB31 (also known as WHRN) cDNA through the inner ear successfully infected numerous HCs, where the localization of whirlin was observed. In the cells where whirlin was expressed, an elongation of the stereocilia bundles was initiated, resulting in the restoration of stereocilia to a wild-type length, which is critical for IHCs to be functional. This whirlin gene therapy approach was able to successfully restore balance and auditory function in Whirler mice for approximately four months [72].
Recombinant AAV vectors, capsid serotype 2 or 8, containing Clrn1 cDNA and driven by a small chicken β-actin (smCBA) promoter, were constructed and injected directly into the inner ear through the RWM of KO-TgAC1 mice to restore the Clrn1 expression [73]. Following a single in vivo injection of the AAV2/8-Clrn1-UTR vector into the cochlea on P1–P3, the KO-TgAC1 mice displayed normal hearing through to adult life and the preservation of the hair bundle structure. Transfection with AAV-Clrn1-UTR successfully suppressed the progressive HL phenotype displayed in KO-TgAC1 mice, offering a potential new gene therapy for treating USH3A [73].
Ush1g knockout mice, which lacked Sans, displayed no ABR to sound, indicating complete deafness, as well as severe vestibular dysfunction [74]. A single dose of AAV8 containing a replacement Ush1g gene delivered locally into the inner ear of newborn knockout mice completely restored the vestibular function to wild-type levels and restored hearing of low frequency loud sounds (≥75 dB) [74]. Gene replacement therapy is quite promising; however, the applicability of this for HL treatment is still challenging due to the characteristics and anatomy of the inner ear structure [75].
Akouos, Inc. has several ongoing development programs involving AAV development and gene therapy. The most advanced program is AK-OTOF, which aims to deliver the OTOF gene to cochlear HCs in patients with hereditary SNHL resulting from mutations in the aforementioned gene [48]. AK-OTOF is currently undergoing Phase I–II clinical trials. Other gene therapies developed by Akouos, Inc. (Boston, MA, USA), including CLRN1 and GJB2, are in preclinical development (Table 2).
A current limitation of AAVs is their transfection efficiency in vivo. While AAVs are able to cross the RWM, their ability to reach the cochlea may be hampered by anti-AAV antibodies and clearance via the cochlear aqueduct [76]. György et al. [77] designed an approach to increase transfection efficiency using exosomes to encapsulate AAV1, which resulted in increased transfection efficiency in vivo compared to AAV1 alone and a partial rescue of hearing in Lhfpl5−/− mice. Moreover, off-target effects and toxicity in the central nervous system are possible by passive transport from the cochlear aqueduct [55,78].
Non-viral vectors are an alternative gene delivery method to viral vectors and largely refer to nanoparticles. Nanoparticles are small and advantageous as drug carriers or gene vectors due to their ability to be modified and functionalized with various excipients, such as lipids or polymers like PEG (polyethylene glycol), PLGA (poly(lactic-co-glycolic acid)) and alginate among others, to improve permeability and targeting. Their main advantage over AAVs is that they can be formulated to have larger carrying capacities, allowing them to carry and deliver larger genes and even proteins. Gene transfection with nanoparticles tends to be less durable than with AAV, although this may not be a disadvantage where the prolonged expression of an exogenous gene may cause unwanted effects [79]. We and others have recently reviewed how nanoparticles are formulated for various therapeutic applications to the inner ear, which we would recommend to the interested reader as that is beyond the scope of this review [57,80]. Briefly, nanoparticles can largely be categorized as lipidic or inorganic, with the former being the more established for general nanomedicine. There are general limitations with nanoparticles in their variable biodegradability, consistency of manufacture and less durable transfection than AAVs. However, an ideal formulation and delivery mode is yet to be developed and recent innovations in nanotechnology and drug delivery put an ideal nanoparticle for inner ear drug delivery within reach. In the following sections on CRISPR and RNAi, we discuss some preclinical studies that have employed lipid nanoparticles for drug delivery.

5. Gene Editing Therapies

5.1. CRISPR-Cas Therapeutics

The CRISPR-Cas9 method of genome editing has great potential to be used for treating certain forms of SNHL. CRISPR, or Clustered Regularly Interspaced Palindromic Repeats, is currently the most prominent method for gene editing. The CRISPR system has two components: a Cas nuclease and guide RNA (gRNA), which guides the nuclease to the target area in the genome [81]. The target area is required to be adjacent to a protospacer-adjacent motif (PAM), a short 3–6 bp sequence found throughout the genome. However, recent variants have been identified that recognize much wider PAM sequence varieties and have been dubbed “near PAM-less” [82,83]. Once guided to the target area, the Cas nuclease causes a double-stranded break in the DNA, where modifications can be made alongside the repair of the break [84,85].
CRISPR-mediated genome editing is especially applicable for certain genetic hearing disorders. Many genetic hearing disorders arise from single mutations in the DNA sequence of a gene, which can theoretically be rectified using CRISPR. As the target cell types in the cochlea are terminally differentiated, many of the strategies described later involve disrupting the target allele of an autosomal dominant disorder or using base editing to modify and repair single nucleotides. Like other methods of repair, notably homology-directed repair, this requires cell division to occur to introduce the modification [86]. At this stage, CRISPR-mediated therapy for some forms of genetic HL is in early development; however, more work has recently been produced in this area, which we have summarized in Table 3.
Gao et al. [87] used a CRISPR approach to disrupt the Tmc1 mutation in the Bth mouse model of autosomal dominant HL. The study employed Streptococcus pyogenes Cas9 (SpCas9, one of the most widely used experimentally) to target and disrupt the mutant Tmc1Bth allele in neonatal mice, using liposomes to deliver the gRNA-CRISPR-Cas9 system into the mouse scala media. The CRISPR-mediated therapy mitigated the progressive cochlear HC death exhibited by the Bth model compared to un-injected mice. The treated Bth mice also exhibited improved low–middle frequency ABR thresholds and an increased startle response compared to the un-injected mice [87]. A more recent study by György et al. [88] used a PAM variant of the Staphylococcus aureus Cas9 (SaCas9-KKH), which was able to target the Bth mutation in the mice and the human p.M418K mutation in human cell lines. The effects of this treatment were sustained for longer than the Gao et al. [88] study. Finally, Zheng et al. [89] used an RNAi approach with CRISPR-CasRx, which is an attractive approach as it has a decreased risk of off-target effects, with transcripts being targeted rather than the genome, resulting in improved hearing up to 15 weeks post-treatment. Similar studies have been conducted using SpCas9 and SaCas9-KKH on mouse models of KCNQ4-associated autosomal dominant hearing loss [90,91].
A strategy for gene editing by homology-mediated end joining (HMEJ) was recently presented by Gu et al. [92]. A mutation in the Klhl18 (Kelch-like family member 18) gene led to the dysfunction of IHCs, which resulted in abnormal ABR but normal DPOAE thresholds [93]. AAV9 and AAV-PHP.eB were used to package the CRISPR-Cas9 knock-in system for delivery into the Klhl18lowf mice, which modeled the autosomal recessive condition in humans. Using the HMEJ strategy, they were able to rescue IHC stereocilia morphology eight weeks after the injection of the dual AAV system and demonstrated improved auditory function up to six months post-treatment [92].

5.2. Base Editing Therapeutics

While the sequence-targeting ability of the Cas nuclease is highly useful, double-stranded breaks are not usually practical for the correction of mutations, especially for single base point mutations where the swapping of one nucleotide pair is sufficient. Double-stranded breaks induced by Cas additionally induce DNA repair mechanisms that may either create insertion/deletion polymorphisms (indels), leading to off-target effects, or the reintroduction of the original pathologic variant [94,95]. The most notable innovation in CRISPR technology for overcoming this is the introduction of base editors, which have recently been tested in preclinical models of genetic SNHL. Base editors are able to modify individual base pairs in double-stranded DNA without causing a double-stranded break. They are chimeric proteins consisting of a modified Cas nuclease, where one of the catalytic domains is inactivated, and a single-stranded DNA deaminase enzyme [84,96]. Base editors are advantageous for gene editing HCs and SCs as mitosis is not required to complete changes to the genome [97]. There are two main classes of base editors: cytosine base editors, which convert C•G base pairs to A•T, and adenosine base editors, which catalyze the inverse A•T to C•G conversion.
Two papers by Yeh and colleagues were the first to demonstrate the potential of base editing in inner ear therapy. The first sought to modify the β-catenin protein to prevent its degradation by the proteasome. Β-catenin is a key effector protein in the Wnt signaling pathway, a crucial pathway regulating HC development, and is targeted for degradation upon the phosphorylation by GSK3β. The study by Yeh et al. [98] showed that the modification of β-catenin prevented phosphorylation and degradation and prolonged Wnt signaling in vitro. Moreover, the authors used a lipid-mediated approach for delivery, as in Gao et al. [87], and showed specific gene editing in SCs when the base editor was applied in vivo. The second paper showed some improvement in hearing when a base editing approach was applied to treat Baringo mice, which had an autosomal recessive form of HL caused by a T•A-to-C•G mutation leading to a non-functional TMC1 protein [99]. This particular approach additionally highlighted the advantage of base editing in treating autosomal recessive disorders by being able to specifically correct the mutation in both alleles in terminally differentiated cells.
Very recently, base editing was applied to an animal model of USH type 1 [100]. This study marked the first in vivo application of CRISPR for USH. The authors generated a mouse model of USH1F, first using CRISPR to introduce the R245X (c.733C•T) mutation, which produced a stop codon in place of an arginine, resulting in a truncated Pcdh15 protein and disrupted hair bundles. The authors used a dual AAV vector with a split intein system to deliver adenine base editors to the cochlea of Pcdh15R245X/R245X mice at P0–P1 and a late-stage conditional knockout model where Pcdh15 was deleted at P15.5 under the control of the Myo15-Cre promotor. The adenine base editors were successful in correcting the C•T mutation for A•G in vivo. However, they were unable to restore hearing in the Pcdh15R245X/R245X mice [100]. The base editor was able to restore hearing in the late-stage conditional knockout model, indicating that the therapy may potentially be more successful if administered in the prenatal period.

5.3. CRISPR in Human Models of SNHL

While animal models have been useful for disease modeling and the functional evaluation of therapeutics, human phenotypes are not always replicated, usually due to polymorphisms of protein-coding genes. This is particularly relevant for assessing the effect of CRISPR on the editing of causative genes where certain variants may lead to different phenotypes in animals and humans. As human tissue samples are scarce, in vitro cell-based models are used for convenience and regenerative capacity and are sufficient for investigating gene editing at the molecular level. Patient-derived cell models have the additional benefits of being relatively easy and non-invasive to establish and numerous studies have reprogrammed various patient cell types into induced pluripotent stem cells (iPSCs) [101,102,103,104,105].
There are numerous studies that have generated cell models from patients with various subtypes of USH-associated genes and used CRISPR to correct the mutation, which we have summarized in Table 3. A notable study by Tang et al. [101] investigated the use of CRISPR-Cas9 to correct the MYO7A mutation in cells derived from a patient with profound HL. The authors generated three iPSC lines. The first cell line was a compound heterozygous mutant from the patient with two MYO7A mutations, c.1184G>A and c.4118C>T, and the second was generated from the asymptomatic father of the patient with only the MYO7A c.1184G>A mutation. The final cell line was derived from a normal donor. The authors used CRISPR-Cas9 and a HDR template to correct the c.4118C>T mutation in the iPSCs and then differentiate them into HC-like cells. A comparison of the corrected patient cell line showed a similar stereocilia bundle morphology to the control and asymptomatic cells, while the unedited patient cell line had curved, disconnected and shorter stereocilia. The corrected patient cell line additionally displayed a comparable electrophysiological response to the asymptomatic and patient cell lines [101].
Similarly, Sanjurjo-Soriano et al. [105] reported the successful seamless correction of the two most prevalent USH2A mutations (c.2276G>T or c.2299delG) in USH2A patient-derived iPSCs via a CRISPR-Cas9 system. The iPSCs were corrected using the high target efficacy and specificity of eSpCas9, without any off-target mutagenesis identified in the corrected iPSCs. These corrected iPSCs retained their pluripotency characteristics and genetic stability. Furthermore, they identified aberrant mRNA levels associated with both mutations that reverted to wild-type following the CRISPR-Cas9 correction, uncovering a novel mechanism for c.2299delG mutation pathogenesis. Another study demonstrated the successful editing of USH2A patient-derived fibroblasts containing the 2299delG mutation [102].
Human cell models have not been limited to USH. Recently, a study applied CRISPR to patient-derived cells with a mitochondrial 12S rRNA 1555A>G and TRMU c.28G>T allele mutation, which has been associated with non-syndromic HL and an increased susceptibility to aminoglycoside-induced HL [106]. This patient had non-syndromic HL and the authors sought to investigate whether a disruption of the TRMU c.28G>T allele, which interacts with the 12S rRNA 1555A>G mutation, would improve the functional properties in the cell model. The authors generated cell lines from the patient, an asymptomatic family member with only the 12S rRNA 1555A>G mutation and an unrelated control, corrected the TRMU c.28G>T allele using CRISPR-Cas9 and then differentiated the cells into HC-like cells to assess their function. The corrected cells displayed more robust differentiation into HC-like cells with a comparative stereocilia morphology, electrophysiological function, mitochondria count and HC-specific gene expression to the cells of the asymptomatic family member [107].
Another study generated iPSCs from the dermal fibroblasts of a patient with profound HL who had compound heterozygous mutations c.4642G>A and c.8374G>A in the MYO15A gene, which is the third most prevalent congenital HL gene [108]. iPSCs were also generated from their normal hearing father, who had the c.4642G>A mutation, and a healthy donor. These iPSC lines were differentiated into HC-like cells. The patient-derived cell line exhibited severely retarded stereocilia formation after differentiation and eventually formed syncytia or died. CRISPR-Cas9 with a HDR template was used to correct the c.4642G>A mutation in the patient-derived iPSCs, which resulted in restored stereocilia formation and no syncytia being formed [109].
CRISPR-Cas technology is continually being improved with innovations in protein engineering and evolution. The newly established prime-editing technique, for example, improves upon base editing by having the ability to install a greater variety of single base changes in the genome, which is highly applicable to autosomal recessive forms of genetic SNHL. While the existing concerns of off-target genome editing remain, the more clinically relevant concerns are when to introduce CRISPR-Cas therapy. For example, if genome editing intervention at the fetal stage is practicable given the terminal development of the neonatal inner ear, would the ideal delivery vehicle be a viral or nanoparticle vector? Moreover, given the promising applications of CRISPR for treating genetic forms of HL and the existing use of patient-derived cell lines, patient screening and personalized medicine approaches are critical for timely intervention with future CRISPR-based therapies.
Table 3. Therapeutic applications of CRISPR-Cas and related systems for SNHL.
Table 3. Therapeutic applications of CRISPR-Cas and related systems for SNHL.
Gene/Disease ModelModelType of EditDelivery ModeEffectReference
DFNA36Bth mice, P1–P2Allele disruption by SpCas9Lipid nanoparticleDecreased OHC death compared to untreated; improved low–middle frequency hearing[87]
DFNA36Bth mice, P1–P2Allele disruption by SaCas9-KKHAnc80L65Preservation of low–middle frequency hearing and decreased HC death; measured up to 24 weeks[88]
DFNA36Bth mice, P1–P2mRNA transcript disruption by CRISPR-CasRxAnc80L65Improved hearing, HC survival and hair bundle formation; measured up to 15 weeks[89]
KCNQ4Knock-in mouse modelSpCas9Lipid nanoparticle or AAV2/Anc80L65Improved auditory function; OHC death not significantly abrogated; no difference in efficacy with injection routes[91]
KCNQ4KCNQ4G229D mice, P1–P2Allele disruption by SaCas9-KKHAAV-PHP.eBImproved auditory function compared to untreated, increased survival and hyperpolarized resting potentials of OHCs; measured up to 12 weeks[90]
KLHL18Homozygous Klhl18lowf miceGene correction by HMEJAAV9 and AAV-PHP.eBHair bundle morphology rescued and improved auditory function up to 6 months post-treatment[92]
DFNA22Myo6WT/C442Y mice, P0–P2SaCas9-KKHAAV-PHP.eBImproved auditory function up to 5 months post-treatment; increased HC survival and improved hair bundle morphology[110]
MYO7APatient-derived iPSCsGene correction with SpCas9 and HDRN/A in vitro studyEdited cells displayed recovered stereocilia morphology and electrophysiological response when differentiated into HC-like cells compared to the differentiated unedited cells[101]
MYO15APatient-derived iPSCsGene correction with Cas9 and HDRN/A in vitro studyEditing restored hair bundle morphology and electrophysiological responses in the differentiated cells compared to the unedited differentiated cells[109]
Neomycin-induced HLNeomycin-treated miceDisruption of HTRA2 gene by SaCas9AAV2/Anc80L65Inhibition of OHC apoptosis, improvement of ABR thresholds; measured up to 8 weeks after neomycin treatment; 1.73% editing efficiency in vivo[111]
TMC1 autosomal recessiveBaringo mice, P1Cytosine base editing of Tmc1Y182C/Y182CDual Anc80L65 vectorsRestoration of HC bundle morphology, sensory transduction current, partial restoration of hearing with improved ABR threshold but no recovery in DPOAE[99]
TRMUPatient-derived iPSCsGene correction of TRMU c.28G>T with SpCas9 and HDRN/A in vitro studyEdited cells differentiated into HC-like cells more efficiently, with improved stereocilia bundle formation, electrophysiological function, mitochondria count and HC marker expression compared to the unedited cells[107]
Usher syndrome type 1FPcdh15R245X miceAdenine base editing of c.733C>T mutationDual AAV9-PHP.B vectorsCorrected causative c.733C>T substitution mutation in vivo; however, hearing was not restored[100]
Usher syndrome type 2A Patient-derived dermal fibroblastsGene correction with SpCas9 and HDRN/A in vitro studyCorrection of c.2299delG deletion mutation in patient dermal fibroblasts[102]
Usher syndrome type 2APatient-derived iPSCsGene correction with SpCas9 and HDRN/A in vitro studyCorrection of c.2299delG deletion mutation in patient iPSCs[112]
Usher syndrome type 2A Patient-derived iPSCsGene correction with eSpCas9 and HDRN/A in vitro studyCorrection of c.2299delG and c.2276G>T mutations in patient iPSCs[105]
ABR, auditory brainstem response; DPOAE, distortion product otoacoustic emission; HC, hair cell; HDR, homology-directed repair; HMEJ, homology-mediated end joining; iPSCs, induced pluripotent stem cells; OHC, outer hair cell.

6. Antisense Oligonucleotides

Antisense oligonucleotides (ASOs) are short, single-stranded nucleotides complementary in sequence to a particular region of a gene that can alter the way that RNA is transcribed, spliced or translated upon binding. There are currently two approaches of nucleotide-based treatments for hearing disorders, namely splice-switching ASOs and RNA interference (RNAi).

6.1. Splice-Switching ASOs

The correct splicing of pre-mRNA to a mature mRNA strand is vital to produce functional proteins. There are many splicing sites found throughout the genome with highly conserved motifs to ensure the gene expression is regulated, despite the majority not being used [113]. Typically, these well-defined splice sites are recognized by the spliceosome to excise and ligate exons embedded in introns to form mature mRNA. However, mutations that interfere with authentic splicing can cause cryptic splice sites as they create a new splice site that is usually not recognized [114]. To suppress and redirect the aberrant splice site, splice-switching ASOs manipulate the splice site by binding complementarily in an antisense orientation to a specific pre-mRNA sequence and interfering with RNA duplex or normal protein–RNA interactions [115]. Splice-switching ASOs create a steric block on the pre-mRNA strand by inhibiting or altering splicing near the site of the deleterious mutations and without interfering with the endogenous gene expression of other associated genes or regulatory elements.
In early in vivo studies, ASO-based therapeutics were targeted by nucleases and rapidly degraded due to their unmodified phosphate backbone and/or sugar ring of the nucleotide [116,117]. Currently, the chemical structure of ASOs can be modified and specified to the intended target site in the body. Improving the binding affinity, cellular uptake and stability are important considerations for in vivo administration. The most common chemical features include substituting the non-bridging oxygen to a sulfur atom on the phosphorothioate backbone and adding a methyl group on the 2′ sugar position for 2′O-methoxyethyl. These modifications provide stability in vivo and resistance against endogenous Rnase-H nucleases, respectively [115,118].
ASO-based treatments have shown promising therapeutic results in preclinical animal models for inner ear diseases. Ush1c mice exhibited disrupted stereocilia bundles, resembling clinical presentation of USH patients [119,120]. The first study to describe ASO-29 as a therapeutic treatment for Ush1c c.216G>A mutation was described in 2013 in mice where the mutation created a cryptic 5′ splice site in exon 3 of harmonin, resulting in a frameshift mutation producing a truncated protein [121]. ASO-29 blocked the cryptic splice site and promoted authentic splicing that resulted in the full-length Ush1c mRNA being transcribed. Ush1c mice treated with ASO-29 at different postnatal days showed significant improvement in their hearing ability. Interestingly, the P3 and P5-treated mutant mice were able to respond to 8–16 kHz sounds but not the P10-treated mutant mice, implying that early treatment may improve the hearing outcome [122].
As the expression of harmonin begins at E15 in mice and is the highest during HC differentiation from P4–P16 [123], multiple strategies to deliver ASO-29 at various stages of development have been reported. ASO-29 was administered via transuterine microinjection into the amniotic cavity and directly into the otic vesicle at E13-E13.5 and E12.5, respectively [124]. While significant hearing preservation at 8, 16 and 24 kHz and partially at 32 kHz was demonstrated with intra-otic delivery, the sub-optimal restoration of hearing to a certain extent was observed in Ush1c mice treated with an amniotic delivery method. In a related study, the same group corrected the Ush1c.216A mutation in mice with ASO-29 to (i) determine the appropriate ASO dosage and age to rescue auditory HCs and (ii) improve on the Lentz et al. [122] study, in which they observed that treatment must be done at earlier stages to rescue inner ear function [125]. Mutant mice received either one (on P1, P5 or P7), two (at P1 and P3) or four (at P1, P3, P5 and P7) doses of ASO-29 by intraperitoneal injection and all showed the expression of the corrected Ush1c mRNA transcript. The ABRs showed variation in the thresholds when simulating for the IHC response regardless of single or multiple dose ASO-29 treatments across post-treatment examination at 1, 3 and 6 months. However, the study did not detect the OHC functional activity by DPOAE when mutant mice received ASO-29 treatment at P7 [125]. It was, therefore, postulated that the therapeutic window for treating defective harmonin is short and must be achieved before HC development or in the early postnatal period in Ush1c mice [43,125]. Indeed, a recent follow-up study of direct RWM injection of ASO-29 to the Ush1c216AA mice inner ear at P1 showed a significant improvement in the hearing function and hair bundle morphology when compared to middle ear administration at later postnatal stages [126].
Disease-causing deep intronic mutations can also be targeted by ASOs. One such mutation in USH2A (c.7595-2144A>G) creates a splice donor site in intron 40 that causes the insertion of a 152-base-pair pseudoexon (PE40) [127]. This insertion was predicted to result in a premature stop codon (p.Lys2532Thrfs*56) and a truncated usherin protein [127]. This mutation is the second most common USH2A-causing mutation, accounting for 4% of USH2A patients [128]. Two ASOs designed to target the PE40 region have shown promising results in USH2A patient-derived fibroblasts harboring a single copy of this mutation. Both ASOs resulted in a reduction in the PE40-containing transcript and an increase in the correctly spliced transcript. Additionally, the combination treatment with both ASOs had an additive effect and was highly effective at correcting splicing in vitro [128].
Similarly, the full length of SLC26A4 exon 8 was rescued by ASO-1029 treatment in a homozygous patient-derived peripheral blood mononuclear cell line harboring the c.919-2A>G mutation. Feng et al. [129] also reported a successful correction of the mis-splicing mutation in Slc26a4 in a humanized mice model via the exon inclusion method. Recently, a study developed a combination of ASOs and the CRISPR-Cas9 system to target the CLRN1 mutation at the mRNA and genomic DNA level, respectively. Mutations in CLRN1 resulted in USH3A, which is characterized by progressive hearing impairments. The splicing correction of the CLRN1 c.256-649T>G mutation was demonstrated when ASOs or CRISPR-Cas9 were administered alone; however, a cocktail of both treatments did not provide any synergistic effect and rather reduced the efficiency of SpCas9 activity for gene correction [130].

6.2. RNA Interference (RNAi)

RNA interference (RNAi) is an endogenous cellular mechanism for targeting foreign double-stranded RNA, such as those introduced by RNA viruses [131]. Recently, the RNAi pathway has been touted as a potential therapeutic approach for targeting genes of interest due to its high potency and specificity. RNAi is stimulated by small double-stranded RNA molecules, including the small interfering RNAs (siRNA) or microRNAs (miRNA), which are recognized by the pathway and digested into shorter fragments by the RNase III-like enzyme Dicer. The fragments are conjugated with an Argonaute protein, forming the RNA-inducing silencing complex (RISC), which guides one strand to form double-stranded RNA that is targeted for cleavage, degradation or repression, while the other strand is removed [132].
siRNAs are derived from double-stranded RNA that has been processed by Dicer into short strands of 21 to 30 nucleotides in length. siRNAs activate the RISC by recruiting the Argonaute protein, which cleaves the sense strand while the antisense strand is guided and remains bound to the RISC. The antisense strand-RISC complex is targeted for complementary base pairing to target mRNA, forming double-stranded RNA, and is processed to be further cleaved. The miRNA-mediated RNAi pathway differs from the siRNA pathway but also employs the Dicer and Argonaute proteins. To effectively target a gene of interest for degradation, there are a number of design considerations. For example, increasing the length of synthetic siRNAs is crucial to target intended mRNA. This increases specificity to the target transcript. Moreover, two overhanging nucleotides must be left on the 3′ end for RNAi machinery recognition [132].
New strategies have recently been developed to rescue deafness through animal models using RNAi [133]. There are many studies experimenting with synthetic siRNA and miRNA to stimulate the RNAi pathway as a therapeutic approach to induce gene silencing for otoprotection. For example, silencing the genes involved in apoptosis, inflammation and oxidative stress, including TNFα, NOX3 and STAT1, has showed protective effects on HCs and neurons [134,135,136,137].
Moreover, RNAi therapy for autosomal dominant forms of HL could be targeted through the knockdown of the dominant mutant allele, which we have summarized in Table 4. For example, Maeda et al. [138] designed a synthetic siRNA to form a degradable duplex to suppress the dominant R75W mutation in the mouse GJB2 gene that causes deafness. In DFNA9, a form of dominantly inherited HL, the c.151C>T founder mutation in COCH has been shown to cause progressive hearing impairment in the Belgian–Dutch population. The COCH gene encodes for cochlin, a highly expressed protein in fibrocytes of the spiral ligament and spiral limbus. An alternative approach to the RNAi-mediated pathway was investigated in targeting mutant COCH transcripts with ASOs for directed degradation by the RNase-H1 mechanism. This was achieved when delivered to COCH minigene constructs. Their most potent ASO was able to induce a 60% reduction in mutant COCH transcripts without affecting the wild-type COCH mRNA levels [139].
Recently, an miRNA-based approach, termed miTmc, was used to suppress the mutant allele in Bth mice. Bth mice received a single trans-RWM injection of either miTmc or miSafe (control treatment) contained in the AAV vector at P0–P2. The mutant mice showed promising results through the significant preservation of the hearing function and improved HC survival for up to 21 weeks after miTmc treatment. The stereocilia bundle formation in miTmc-treated Bth mice was preserved when compared to the untreated and miSafe-treated Bth mice where severe degeneration was observed [140]. A study on miRNA treatment in Bth mice by Yoshimura et al. [141] administered a single dose of miTmc to knock down the mutant allele at P15–P16 and was sufficient to slow the progression of HL. Moreover, they administered miTmc in Bth mice at two additional timepoints, P56–P60 and P84–P90, to test the optimal therapeutic window for rescuing hearing. Interestingly, a mild protective effect on hearing was observed for those treated at P56–P60 and, as expected, treatment at P84–P90 failed to prevent the deterioration of hearing.
An alternative strategy to regenerate HCs from SCs was proposed through gene silencing of the Hes1/Hes5 gene complex using siRNA-based antisense therapy [142,143]. Notch signaling is essential for cell fate and the specification of sensory progenitor cells in the inner ear. The activation of Notch signaling initiated by ligand-receptor binding promotes the expression of Hes1/Hes5, which induces a SC fate by downregulating Atoh1 and other prosensory genes [144]. Moreover, Hes1 is consistently expressed in adulthood to maintain the appropriate HC and SC organization in the organ of Corti. Du et al. [142] observed supernumerary IHCs localized in the middle turn of the cochlea by silencing the Hes1 and Hes5 genes. Ultimately, the number of p27Kip1-positive SCs was reduced, suggesting that the increase in the HC population was compensated by the number of existing SCs.

7. Cell Therapy

Stem cells are precursor cells defined by their ability to self-renew, i.e., give rise to new stem cells and differentiate into many cell types. These properties make them an attractive therapeutic option for a variety of diseases, including SNHL. There are several types of stem cells, but the three types commonly used in regenerative studies are: (i) embryonic stem cells (ESCs), (ii) adult stem cells (ASCs) and (iii) induced pluripotent stem cells (iPSCs).
iPSCs reprogrammed from somatic cells were first reprogrammed from mouse embryonic and adult fibroblast cultures in 2006 through treatment with specific transcription factors [145]. These pluripotency-related transcription factors, also known as ‘Yamanaka Factors’ or ‘OSKMs’ (named OCT4, SOX2, c-MYC and KLF4), are sufficient to induce pluripotency in different adult cell types [145,146]. Although the roles of c-MYC and KLF4 remain unclear, OCT4 and SOX2 are crucial for activating and maintaining the genes associated with pluripotency and renewal, while simultaneously repressing differentiation-associated genes [147,148]. KLF4 is thought to promote cell survival by exerting an anti-apoptotic function, while c-MYC is thought to be involved in the regulation of histone acetylation. c-MYC was initially considered to be non-essential in the reprogramming process, but it is now thought that c-MYC and KLF4 work together to maintain the immortality of iPSCs [149,150].
There are several published protocols to induce HC regeneration from stem cells, each with varied success. Li and colleagues were the first to publish a protocol on generating inner ear progenitor cells derived from mouse ESCs, through the expression of HC markers shown by immunostaining and gene expression analysis [151]. Based on the principles of early otic development, a more sophisticated in vitro feeder layer-based protocol involving several growth factors was shown to successfully promote the differentiation of mouse ESCs and iPSCs into mechanosensitive HC-like cells that were responsive to electrophysiological stimulation [152].
Stem/progenitor cell transplantation therapy uses “corrected” or healthy cells derived from patients to replace older or damaged cells. Stem cell transplantation therapy is a favorable approach for personalized medicine as adult skin cells can be non-invasively harvested from patients and reprogrammed into a pluripotent state. This method of therapeutic approach minimizes incompatibility risks of donor/recipient transplants when administered to patients.
Some studies have reported the use of ASCs/inner ear progenitor cells to form HCs or neural-like cells. Chen et al. [153] transplanted otic-like neural progenitor cells into a mouse model of auditory neuropathy and demonstrated significantly improved ABR thresholds by 46% compared to the untreated mice. In a similar study, Barboza et al. [154] used mouse inner ear progenitor stem cells for HC replacement in adult guinea pigs treated with neomycin. Differentiated cells expressing HC marker MYO7A and SC marker p27Kip1 were generated from mature organ of Corti-derived cells and transplanted into the inner ear of the neomycin-treated guinea pigs. The transplanted cells localized in the cochlear basilar membrane and improved the auditory thresholds of the treated guinea pigs.
Similarly, Chen et al. [155] transplanted human iPSC-derived HC-like cells into Slc26a4-null mice. Urinary cells from a healthy donor were reprogrammed into iPSCs. These human iPSCs were differentiated into HC-like cells and, separately, SGN-like cells, confirmed by the cell fate markers and electrophysiology studies. When co-cultured, the HC-like cells were able to form synapses with mouse SGNs and differentiated SGN-like cells, as indicated by the expression of CtBP2 and SYP. When HC-like cells were injected into the RWM of Slc26a4-null mice, a small population were able to migrate to the organ of Corti, implant into the native cells and maintain the MYO7A expression. The native SGNs were able to form synapses with the implanted cells. The implanted HC-like cells demonstrated no risk of tumorigenesis when injected into NOD/SCID mice. However, the HC-like cells failed to implant in the regular organization of native HCs and hearing was not assessed in this study [155].
In another study, stem cell transplantation was used to treat mice with a connexin 30 (Cx30) deletion. Healthy human iPSC-derived outer sulcus-like cells were engrafted into E11.5 mouse cochlea to test their capacity for stem cell regeneration. Interestingly, immunohistochemistry staining of the treated cochleae showed that the engrafted cells expressed Cx30 one week after transplantation [156]. Although Cx30-positive cells were detectable in the non-sensory region in the cochlea, only a limited number of engrafted cells were observed in the inner ear. Similar to the study by Lee et al. [157], only 1–3% of the engrafted cells survived.
A study by Takeda et al. [158] investigated whether the ablation of native HCs would improve the engraftment of transplanted HCs into the organ of Corti. Ablation was achieved by treating Pou4f3DTR/+ mice with diphtheria toxin, which specifically deleted HCs. Human ESCs at day seven of otic differentiation were determined to be at the pre-placodal stage and injected in the scala tympani perilymph via the RWM of neonatal Pou4f3DTR/+ mice treated with diphtheria toxin to specifically ablate IHCs. While the treated Pou4f3DTR/+ mice showed a higher migration and engraftment of HC-like cells in the organ of Corti, the quantity of cells successfully engrafted was still low and the mice did not show improvement in the hearing thresholds.
Adult tissue-specific stem cells reside in somatic tissues and can self-renew and differentiate into specialized cell types. Mesenchymal stem cells (MSCs) are adult stromal cells that display multipotency with limited self-renewal and differentiation capacity. Moreover, studies conducted in animals have shown that MSCs can be differentiated into auditory HCs and neural cells [159,160,161,162]. MSCs release a wide spectrum of proteins, growth factors, cytokines and unique extracellular vesicles (EVs) with immunomodulatory properties. These EVs have been shown to have a potential therapeutic role in rescuing cochlear damage [163]. A recent study demonstrated that these EVs promoted the activation of SOD2, an oxidoreductase, to maintain the levels of mitochondrial reactive oxygen species and increase resistance to oxidative damage in a mouse model of NIHL [164]. In a clinical report, a patient received cochlear implants and the left ear also received allogenic umbilical cord-derived MSC-EVs, the purpose of which was to reduce inflammation and improve cochlear health during the implant procedure [165]. The patient did not show any signs of toxicity or adverse reaction in follow-up appointments at 24 months and demonstrated a significant improvement in speech perception compared to the untreated ear.
Using autologous cells from patients reduces the risk of cell rejection by the immune system and is another promising cell therapy approach for HC regeneration. This form of transplantation has already been FDA-approved and is currently in use for other forms of neurodegenerative diseases like amyotrophic lateral sclerosis [166]. Recently, a Phase I/II clinical trial aimed at treating acquired HL in children with autologous MSCs derived from human umbilical cord blood (hUCB) for HC regeneration has been completed (NCT02038972). A single infusion of autologous hUCB in children aged 5 months to 5 years old all demonstrated an improvement in the ABR thresholds by 5 dB. A follow-up of post-infusion treatment showed that 10 out of 11 subjects showed an improvement in verbal language scoring [167].
Two pilot clinical studies by Lee et al. [168] aimed to transplant autologous MSCs in SNHL patients with auditory neuropathy. Two patients aged 55 and 67 years received an infusion of bone-marrow-derived MSCs for three years. No systematic complication was reported but there was no sign of any hearing improvement in either case [168]. These results obtained from human studies showed that cell-based therapy is possible for treating HL but due to the limited evidence, further study will be needed before transitioning into clinical use.
A major obstacle for cell replacement therapy is evading the immune response when grafting cells. One strategy for overcoming this was presented by Andrade da Silva et al. [169], who used genetic modification in human iPSCs. The human leukocyte antigen genes, which have the most significant effect on graft rejection, were knocked out using CRISPR-Cas9 to reduce immunogenicity when delivering two dimensional-differentiated otic neural progenitor cells into immunocompetent mouse cochlea. This resulted in increased survival of the edited human iPSCs following intracochlear transplantation compared to the unedited cells in a mouse model.
Although stem cell-based therapy is on the cutting edge of modern medicine, there are several challenges and risks to mitigate before clinical application. The major concerns relate to the shared features of stem cells and cancer cells, such as the ability to self-renew, the potency of stem cells (pluripotent or multipotent), indefinite growth and a high proliferation rate. The transplanted cells would need to survive and grow after their insertion into the patient cochlea, without interfering with normal hearing. Additional risks such as the formation of tumors, potential stem cell migration to unwanted sites and immune rejection of transplanted stem cells must also be considered. Furthermore, there are a number of factors that would require highly rigorous control and oversight, including the type of stem cells used, their procurement and culturing history, the level of manipulation and administration site associated with stem cell-based therapy and medicinal products.

8. Conclusions

This review has provided an overview of therapeutic strategies for the potential treatment of HL. These included gene-based and cell-based therapies in which, as highlighted, the majority in development are focused on preventing and treating HL. Upon the identification of HL, genetic testing may reveal whether a mutation is present in the known causative genes, which also provides information on the natural history and potential systemic associations. It follows, then, that therapeutics could be delivered to a patient before their hearing deteriorates, and maintaining important protein structures and complexes is significantly easier than restoring them.
Unfortunately, for most patients with inherited HL, severe hearing impairment is evident within the first few years of life. Therefore, an essential consideration is the therapeutic window. To have the most therapeutic benefit, the drug would need to be delivered before the causative mutation has compromised the hearing structures, which is impractical in most cases. However, gene therapy trials in animal models have shown some evidence of functional hearing restoration [73,75]. While this has not yet been assessed in human patients, there is definitely hope that these therapies may have a similar restorative effect. Nevertheless, it will be preferable to deliver biological therapies as early as possible to ensure the best possible outcome for patients.
Conversely for patients who have extensive cell death within their organ of Corti, regenerative and cell therapies would be highly applicable to replace HCs and neurons. Cell therapy is becoming increasingly closer to reality as demonstrated by the advances reported in this review. However, the key question of how implanted cells are seeded into the organ of Corti and how some cells survive while most perish remains unanswered. Understanding this mechanism, perhaps starting with a suitable in vitro scaffold modeling the organ of Corti, could help improve the efficacy and survival of cell supplementation therapies for HL. Nevertheless, these seminal studies into inner ear cell therapies have already illuminated a pathway for the eventual replacement of sensory cells into the human cochlea.

Author Contributions

Conceptualization, E.Y.M.W.; writing—original draft preparation, J.L., D.S.D., F.K.L., V.S.S. and E.Y.M.W.; writing—review and editing, J.L., D.S.D., F.K.L., F.K.C. and E.Y.M.W.; supervision, S.M., F.K.C., M.D.A. and E.Y.M.W.; funding acquisition, S.M., F.K.C., M.D.A. and E.Y.M.W. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Australian government National Health and Medical Research Council (NHMRC) Medical Research Future Fund (MRFF) Stem Cell Therapies Mission Grant: AAP2017281, Western Australian (WA) Future Health Research and Innovation Fund (FHRIF) Near-miss Awards: Ideas Grants (WANMA): WANMA/IG2021/6, Commonwealth Scientific and Industrial Research Organization (CSIRO) ON Prime Innovation Program: P12-1426, Passe and Williams Foundation, Perron Charitable Foundation, Channel 7 Telethon grant and WA Department of Jobs, Tourism, Science and Innovative (JTSI) Industries Fund 2023 Innovation Booster Grant. J.L. was supported by a PhD scholarship provided by Ear Science Institute Australia, enrolled through the University of Western Australia. E.Y.M.W. was supported by an Adjunct Associate Professor appointment at the University of Western Australia.

Acknowledgments

The authors thank Peter Millington for his additional review and proof reading of this manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chadha, S.; Kamenov, K.; Cieza, A. The world report on hearing, 2021. Bull World Health Organ. 2021, 99, 242–242A. [Google Scholar] [CrossRef] [PubMed]
  2. Mohr, P.E.; Feldman, J.J.; Dunbar, J.L.; McConkey-Robbins, A.; Niparko, J.K.; Rittenhouse, R.K.; Skinner, M.W. The Societal Costs of Severe to Profound Hearing Loss in The United States. Int. J. Technol. Assess. Health Care 2001, 16, 1120–1135. [Google Scholar] [CrossRef] [PubMed]
  3. Deniz, B.; Boz, C.; Kara, E.; Deniz, R.; Oruç, Y.; Acar, M.; Yılmaz, Y.Z.; Ataş, A. Direct Health Expenditure Analysis Related to Hearing Loss in Individuals Using Hearing Aids and Cochlear Implants. Turk. Arch. Otorhinolaryngol. 2022, 60, 142–148. [Google Scholar] [CrossRef] [PubMed]
  4. Ciorba, A.; Bianchini, C.; Pelucchi, S.; Pastore, A. The impact of hearing loss on the quality of life of elderly adults. Clin. Interv. Aging 2012, 7, 159–163. [Google Scholar] [CrossRef] [PubMed]
  5. Idstad, M.; Tambs, K.; Aarhus, L.; Engdahl, B.L. Childhood sensorineural hearing loss and adult mental health up to 43 years later: Results from the HUNT study. BMC Public. Health 2019, 19, 168. [Google Scholar] [CrossRef] [PubMed]
  6. Bisogno, A.; Scarpa, A.; Di Girolamo, S.; De Luca, P.; Cassandro, C.; Viola, P.; Ricciardiello, F.; Greco, A.; Vincentiis, M.; Ralli, M.; et al. Hearing Loss and Cognitive Impairment: Epidemiology, Common Pathophysiological Findings, and Treatment Considerations. Life 2021, 11, 1102. [Google Scholar] [CrossRef] [PubMed]
  7. 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]
  8. Livingston, G.; Huntley, J.; Sommerlad, A.; Ames, D.; Ballard, C.; Banerjee, S.; Brayne, C.; Burns, A.; Cohen-Mansfield, J.; Cooper, C.; et al. Dementia prevention, intervention, and care: 2020 report of the Lancet Commission. Lancet 2020, 396, 413–446. [Google Scholar] [CrossRef]
  9. Ji, L.; Lee, H.-J.; Wan, G.; Wang, G.-P.; Zhang, L.; Sajjakulnukit, P.; Schacht, J.; Lyssiotis, C.A.; Corfas, G. Auditory metabolomics, an approach to identify acute molecular effects of noise trauma. Sci. Rep. 2019, 9, 9273. [Google Scholar] [CrossRef]
  10. Ashmore, J.; Gale, J. The cochlea. Curr. Biol. 2000, 10, R325–R327. [Google Scholar] [CrossRef]
  11. Hakizimana, P.; Fridberger, A. Inner hair cell stereocilia are embedded in the tectorial membrane. Nat. Commun. 2021, 12, 2604. [Google Scholar] [CrossRef] [PubMed]
  12. Lv, P.; Sihn, C.-R.; Wang, W.; Shen, H.; Kim, H.J.; Rocha-Sanchez, S.M.; Yamoah, E.N. Posthearing Ca2+ Currents and Their Roles in Shaping the Different Modes of Firing of Spiral Ganglion Neurons. J. Neurosci. 2012, 32, 16314. [Google Scholar] [CrossRef]
  13. Neng, L.; Zhang, F.; Kachelmeier, A.; Shi, X. Endothelial Cell, Pericyte, and Perivascular Resident Macrophage-Type Melanocyte Interactions Regulate Cochlear Intrastrial Fluid–Blood Barrier Permeability. JARO J. Assoc. Res. Otolaryngol. 2013, 14, 175–185. [Google Scholar] [CrossRef] [PubMed]
  14. Cosgrove, D.; Zallocchi, M. Usher protein functions in hair cells and photoreceptors. Int. J. Biochem. Cell Biol. 2014, 46, 80–89. [Google Scholar] [CrossRef] [PubMed]
  15. Warren, R.L.; Ramamoorthy, S.; Ciganović, N.; Zhang, Y.; Wilson, T.M.; Petrie, T.; Wang, R.K.; Jacques, S.L.; Reichenbach, T.; Nuttall, A.L.; et al. Minimal basilar membrane motion in low-frequency hearing. Proc. Natl. Acad. Sci. USA 2016, 113, E4304–E4310. [Google Scholar] [CrossRef] [PubMed]
  16. Cunningham, L.L.; Tucci, D.L. Hearing Loss in Adults. N. Engl. J. Med. 2017, 377, 2465–2473. [Google Scholar] [CrossRef] [PubMed]
  17. Rizk, H.G.; Lee, J.A.; Liu, Y.F.; Endriukaitis, L.; Isaac, J.L.; Bullington, W.M. Drug-Induced Ototoxicity: A Comprehensive Review and Reference Guide. Pharmacotherapy 2020, 40, 1265–1275. [Google Scholar] [CrossRef]
  18. Kakarlapudi, V.; Sawyer, R.; Staecker, H. The Effect of Diabetes on Sensorineural Hearing Loss. Otol. Neurotol. 2003, 24, 382–386. [Google Scholar] [CrossRef]
  19. Kumar, A.; Gulati, R.; Singhal, S.; Hasan, A.; Khan, A. The effect of smoking on the hearing status-a hospital based study. J. Clin. Diagn. Res. 2013, 7, 210–214. [Google Scholar] [CrossRef]
  20. Sterling, M.R.; Lin, F.R.; Jannat-Khah, D.P.; Goman, A.M.; Echeverria, S.E.; Safford, M.M. Hearing Loss Among Older Adults with Heart Failure in the United States: Data From the National Health and Nutrition Examination Survey. JAMA Otolaryngol. Head. Neck Surg. 2018, 144, 273–275. [Google Scholar] [CrossRef]
  21. Tan, H.E.; Lan, N.S.R.; Knuiman, M.W.; Divitini, M.L.; Swanepoel, D.W.; Hunter, M.; Brennan-Jones, C.G.; Hung, J.; Eikelboom, R.H.; Santa Maria, P.L. Associations between cardiovascular disease and its risk factors with hearing loss—A cross-sectional analysis. Clin. Otolaryngol. 2018, 43, 172–181. [Google Scholar] [CrossRef] [PubMed]
  22. Schrijver, I. Hereditary non-syndromic sensorineural hearing loss: Transforming silence to sound. J. Mol. Diagn. 2004, 6, 275–284. [Google Scholar] [CrossRef] [PubMed]
  23. Gettelfinger, J.D.; Dahl, J.P. Syndromic Hearing Loss: A Brief Review of Common Presentations and Genetics. J. Pediatr. Genet. 2018, 7, 001–008. [Google Scholar] [CrossRef] [PubMed]
  24. Kenneson, A.; Van Naarden Braun, K.; Boyle, C. GJB2 (connexin 26) variants and nonsyndromic sensorineural hearing loss: A HuGE review. Genet. Med. 2002, 4, 258–274. [Google Scholar] [CrossRef] [PubMed]
  25. Allen, S.B.; Goldman, J. Syndromic Sensorineural Hearing Loss. Available online: https://www.ncbi.nlm.nih.gov/books/NBK526088/ (accessed on 27 September 2022).
  26. Géléoc, G.S.G.; Holt, J.R. Sound Strategies for Hearing Restoration. Science 2014, 344, 1241062. [Google Scholar] [CrossRef] [PubMed]
  27. Mirian, C.; Ovesen, T. Intratympanic vs Systemic Corticosteroids in First-line Treatment of Idiopathic Sudden Sensorineural Hearing Loss: A Systematic Review and Meta-analysis. JAMA Otolaryngol. Head. Neck Surg. 2020, 146, 421–428. [Google Scholar] [CrossRef]
  28. Plontke, S.K.; Meisner, C.; Agrawal, S.; Cayé-Thomasen, P.; Galbraith, K.; Mikulec, A.A.; Parnes, L.S.; Premakumar, Y.; Reiber, J.; Schilder, A.G.M.; et al. Intratympanic corticosteroids for sudden sensorineural hearing loss. Cochrane Database Syst. Rev. 2022. [Google Scholar] [CrossRef]
  29. Salt, A.N.; Plontke, S.K. Pharmacokinetic principles in the inner ear: Influence of drug properties on intratympanic applications. Hear. Res. 2018, 368, 28–40. [Google Scholar] [CrossRef]
  30. 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. Eng. J. Med. 2018, 378, 2376–2385. [Google Scholar] [CrossRef]
  31. Sooriyaarachchi, M.; George, G.N.; Pickering, I.J.; Narendran, A.; Gailer, J. Tuning the metabolism of the anticancer drug cisplatin with chemoprotective agents to improve its safety and efficacy. Met. Integr. Biometal Sci. 2016, 8, 1170–1176. [Google Scholar] [CrossRef]
  32. Jiam, N.T.; Rauch, S.D. Inner ear therapeutics and the war on hearing loss: Systemic barriers to success. Front. Neurosci. 2023, 17, 1169122. [Google Scholar] [CrossRef] [PubMed]
  33. Driver, E.C.; Kelley, M.W. Development of the cochlea. Development 2020, 147, dev162263. [Google Scholar] [CrossRef] [PubMed]
  34. Koehler, K.R.; Nie, J.; Longworth-Mills, E.; Liu, X.-P.; Lee, J.; Holt, J.R.; Hashino, E. Generation of inner ear organoids containing functional hair cells from human pluripotent stem cells. Nat. Biotechnol. 2017, 35, 583–589. [Google Scholar] [CrossRef] [PubMed]
  35. Koehler, K.R.; Mikosz, A.M.; Molosh, A.I.; Patel, D.; Hashino, E. Generation of inner ear sensory epithelia from pluripotent stem cells in 3D culture. Nature 2013, 500, 217–221. [Google Scholar] [CrossRef] [PubMed]
  36. Jeong, M.; O’Reilly, M.; Kirkwood, N.K.; Al-Aama, J.; Lako, M.; Kros, C.J.; Armstrong, L. Generating inner ear organoids containing putative cochlear hair cells from human pluripotent stem cells. Cell Death Dis. 2018, 9, 922. [Google Scholar] [CrossRef] [PubMed]
  37. Quan, Y.-Z.; Wei, W.; Ergin, V.; Rameshbabu, A.P.; Huang, M.; Tian, C.; Saladi, S.V.; Indzhykulian, A.A.; Chen, Z.-Y. Reprogramming by drug-like molecules leads to regeneration of cochlear hair cell–like cells in adult mice. Proc. Natl. Acad. Sci. USA 2023, 120, e2215253120. [Google Scholar] [CrossRef] [PubMed]
  38. Zhang, X.; Li, Y.; Xu, H.; Zhang, Y.-W. The γ-secretase complex: From structure to function. Front. Cell Neurosci. 2014, 8, 427. [Google Scholar] [CrossRef]
  39. Brown, R.; Groves, A.K. Hear, Hear for Notch: Control of Cell Fates in the Inner Ear by Notch Signaling. Biomolecules 2020, 10, 370. [Google Scholar] [CrossRef]
  40. Polyak, K.; Kato, J.Y.; Solomon, M.J.; Sherr, C.J.; Massague, J.; Roberts, J.M.; Koff, A. p27Kip1, a cyclin-Cdk inhibitor, links transforming growth factor-beta and contact inhibition to cell cycle arrest. Genes. Dev. 1994, 8, 9–22. [Google Scholar] [CrossRef]
  41. Mantela, J.; Jiang, Z.; Ylikoski, J.; Fritzsch, B.; Zacksenhaus, E.; Pirvola, U. The retinoblastoma gene pathway regulates the postmitotic state of hair cells of the mouse inner ear. Development 2005, 132, 2377–2388. [Google Scholar] [CrossRef]
  42. Laine, H.; Doetzlhofer, A.; Mantela, J.; Ylikoski, J.; Laiho, M.; Roussel, M.F.; Segil, N.; Pirvola, U. p19Ink4d and p21Cip1 Collaborate to Maintain the Postmitotic State of Auditory Hair Cells, Their Codeletion Leading to DNA Damage and p53-Mediated Apoptosis. J. Neurosci. 2007, 27, 1434–1444. [Google Scholar] [CrossRef] [PubMed]
  43. Walters, B.J.; Liu, Z.; Crabtree, M.; Coak, E.; Cox, B.C.; Zuo, J. Auditory Hair Cell-Specific Deletion of p27Kip1 in Postnatal Mice Promotes Cell-Autonomous Generation of New Hair Cells and Normal Hearing. J. Neurosci. 2014, 34, 15751–15763. [Google Scholar] [CrossRef] [PubMed]
  44. Chen, P.; Segil, N. p27Kip1 links cell proliferation to morphogenesis in the developing organ of Corti. Development 1999, 126, 1581–1590. [Google Scholar] [CrossRef] [PubMed]
  45. Löwenheim, H.; Furness, D.N.; Kil, J.; Zinn, C.; Gültig, K.; Fero, M.L.; Frost, D.; Gummer, A.W.; Roberts, J.M.; Rubel, E.W.; et al. Gene disruption of p27Kip1 allows cell proliferation in the postnatal and adult organ of Corti. Proc. Natl. Acad. Sci. USA 1999, 96, 4084–4088. [Google Scholar] [CrossRef] [PubMed]
  46. Kil, J.; Gu, R.; Fero, M.; Lynch, E.D. The Cyclin Dependent Kinase Inhibitor p27Kip1 Maintains Terminal Differentiation in the Mouse Organ of Corti. Open Otorhinolaryngol. J. 2011, 5, 25–34. [Google Scholar] [CrossRef]
  47. Jones, D. To hear a whisper: Biotechs chase new thinking to restore hearing. Nat Biotechnol. 2018, 36, 1128–1129. [Google Scholar] [CrossRef]
  48. Sly, D.J.; Campbell, L.; Uschakov, A.; Saief, S.T.; Lam, M.; O’Leary, S.J. Applying Neurotrophins to the Round Window Rescues Auditory Function and Reduces Inner Hair Cell Synaptopathy After Noise-induced Hearing Loss. Otol. Neurotol. 2016, 37, 1223–1230. [Google Scholar] [CrossRef]
  49. Szobota, S.; Mathur, P.D.; Siegel, S.; Black, K.; Saragovi, H.U.; Foster, A.C. BDNF, NT-3 and Trk receptor agonist monoclonal antibodies promote neuron survival, neurite extension, and synapse restoration in rat cochlea ex vivo models relevant for hidden hearing loss. PLoS ONE 2019, 14, e0224022. [Google Scholar] [CrossRef]
  50. Wise, A.K.; Richardson, R.; Hardman, J.; Clark, G.; O’Leary, S. Resprouting and survival of guinea pig cochlear neurons in response to the administration of the neurotrophins brain-derived neurotrophic factor and neurotrophin-3. J. Comp. Neurol. 2005, 487, 147–165. [Google Scholar] [CrossRef]
  51. Pollack, S.J.; Harper, S.J. Small Molecule Trk Receptor Agonists and Other Neurotrophic Factor Mimetics. CNS Neurol. Disord. Drug Targets 2002, 1, 59–80. [Google Scholar] [CrossRef]
  52. Qing, Y.; Qing, C.; Xia, L.; Yunfeng, W.; Huawei, L.; Shusheng, G.; Keqiang, Y.; Xi, L. Protection of Spiral Ganglion Neurons from Degeneration Using Small-Molecule TrkB Receptor Agonists. J. Neurosci. 2013, 33, 13042. [Google Scholar] [CrossRef]
  53. Fernandez, K.A.; Watabe, T.; Tong, M.; Meng, X.; Tani, K.; Kujawa, S.G.; Edge, A.S.B. Trk agonist drugs rescue noise-induced hidden hearing loss. JCI Insight 2021, 6, e142572. [Google Scholar] [CrossRef] [PubMed]
  54. Kempfle, J.S.; Duro, M.V.; Zhang, A.; Amador, C.D.; Kuang, R.; Lu, R.; Kashemirov, B.A.; Edge, A.S.B.; McKenna, C.E.; Jung, D.H. A Novel Small Molecule Neurotrophin-3 Analogue Promotes Inner Ear Neurite Outgrowth and Synaptogenesis In Vitro. Front. Cell Neurosci. 2021, 15, 666706. [Google Scholar] [CrossRef] [PubMed]
  55. Pinyon, J.L.; von Jonquieres, G.; Crawford, E.N.; Duxbury, M.; Al Abed, A.; Lovell, N.H.; Klugmann, M.; Wise, A.K.; Fallon, J.B.; Shepherd, R.K.; et al. Neurotrophin gene augmentation by electrotransfer to improve cochlear implant hearing outcomes. Hear. Res. 2019, 380, 137–149. [Google Scholar] [CrossRef] [PubMed]
  56. Li, L.; Chao, T.; Brant, J.; O’Malley, B.; Tsourkas, A.; Li, D. Advances in nano-based inner ear delivery systems for the treatment of sensorineural hearing loss. Adv. Drug Deliv. Rev. 2017, 108, 2–12. [Google Scholar] [CrossRef] [PubMed]
  57. Valente, F.; Astolfi, L.; Simoni, E.; Danti, S.; Franceschini, V.; Chicca, M.; Martini, A. Nanoparticle drug delivery systems for inner ear therapy: An overview. J. Drug Deliv. Sci. Technol. 2017, 39, 28–35. [Google Scholar] [CrossRef]
  58. Hao, J.; Li, S.K. Inner ear drug delivery: Recent advances, challenges, and perspective. Eur. J. Pharm. Sci. 2019, 126, 82–92. [Google Scholar] [CrossRef]
  59. Xu, X.; Wang, S.; Shu, Y.; Zhang, H. Nanoparticle-based inner ear delivery systems for the treatment of hearing loss. Smart Mater. Med. 2021, 2, 350–353. [Google Scholar] [CrossRef]
  60. Bermingham, N.A.; Hassan, B.A.; Price, S.D.; Vollrath, M.A.; Ben-Arie, N.; Eatock, R.A.; Bellen, H.J.; Lysakowski, A.; Zoghbi, H.Y. Math1 An Essential Gene for the Generation of Inner Ear Hair Cells. Science 1999, 284, 1837–1841. [Google Scholar] [CrossRef]
  61. Ben-Arie, N.; Bellen, H.J.; Armstrong, D.L.; McCall, A.E.; Gordadze, P.R.; Guo, Q.; Matzuk, M.M.; Zoghbi, H.Y. Math1 is essential for genesis of cerebellar granule neurons. Nature 1997, 390, 169–172. [Google Scholar] [CrossRef]
  62. Atkinson, P.J.; Wise, A.K.; Flynn, B.O.; Nayagam, B.A.; Richardson, R.T. Hair Cell Regeneration after ATOH1 Gene Therapy in the Cochlea of Profoundly Deaf Adult Guinea Pigs. PLoS ONE 2014, 9, e102077. [Google Scholar] [CrossRef] [PubMed]
  63. Allocca, M.; Doria, M.; Petrillo, M.; Colella, P.; Garcia-Hoyos, M.; Gibbs, D.; Kim, S.R.; Maguire, A.; Rex, T.S.; Di Vicino, U.; et al. Serotype-dependent packaging of large genes in adeno-associated viral vectors results in effective gene delivery in mice. J. Clin. Inv. 2008, 118, 1955–1964. [Google Scholar] [CrossRef] [PubMed]
  64. Lopes, V.S.; Boye, S.E.; Louie, C.M.; Boye, S.; Dyka, F.; Chiodo, V.; Fofo, H.; Hauswirth, W.W.; Williams, D.S. Retinal gene therapy with a large MYO7A cDNA using adeno-associated virus. Gene Ther. 2013, 20, 824–833. [Google Scholar] [CrossRef] [PubMed]
  65. Dyka, F.M.; Boye, S.L.; Chiodo, V.A.; Hauswirth, W.W.; Boye, S.E. Dual Adeno-Associated Virus Vectors Result in Efficient In Vitro and In Vivo Expression of an Oversized Gene, MYO7A. Hum. Gene Ther. Methods 2014, 25, 166–177. [Google Scholar] [CrossRef] [PubMed]
  66. Trapani, I.; Colella, P.; Sommella, A.; Iodice, C.; Cesi, G.; de Simone, S.; Marrocco, E.; Rossi, S.; Giunti, M.; Palfi, A.; et al. Effective delivery of large genes to the retina by dual AAV vectors. EMBO Mol. Med. 2014, 6, 194–211. [Google Scholar] [CrossRef]
  67. Tertrais, M.; Bouleau, Y.; Emptoz, A.; Belleudy, S.; Sutton, R.B.; Petit, C.; Safieddine, S.; Dulon, D. Viral Transfer of Mini-Otoferlins Partially Restores the Fast Component of Exocytosis and Uncovers Ultrafast Endocytosis in Auditory Hair Cells of Otoferlin Knock-Out Mice. J. Neurosci. 2019, 39, 3394–3411. [Google Scholar] [CrossRef]
  68. Ivanchenko, M.V.; Hathaway, D.M.; Klein, A.J.; Pan, B.; Strelkova, O.; De-la-Torre, P.; Wu, X.; Peters, C.W.; Mulhall, E.M.; Booth, K.T.; et al. Mini-PCDH15 gene therapy rescues hearing in a mouse model of Usher syndrome type 1F. Nat. Commun. 2023, 14, 2400. [Google Scholar] [CrossRef]
  69. Suzuki, S.; Nakajima, T.; Irie, S.; Ariyasu, R.; Komiyama, T.; Ohki, Y. Vestibular stimulation-induced facilitation of cervical premotoneuronal systems in humans. PLoS ONE 2017, 12, e0175131. [Google Scholar] [CrossRef]
  70. Pan, B.; Askew, C.; Galvin, A.; Heman-Ackah, S.; Asai, Y.; Indzhykulian, A.A.; Jodelka, F.M.; Hastings, M.L.; Lentz, J.J.; Vandenberghe, L.H.; et al. Gene therapy restores auditory and vestibular function in a mouse model of Usher syndrome type 1c. Nat. Biotechnol. 2017, 35, 264–272. [Google Scholar] [CrossRef]
  71. Askew, C.; Rochat, C.; Pan, B.; Asai, Y.; Ahmed, H.; Child, E.; Schneider, B.L.; Aebischer, P.; Holt, J.R. Tmc gene therapy restores auditory function in deaf mice. Sci. Transl. Med. 2015, 7, 295ra108. [Google Scholar] [CrossRef]
  72. Isgrig, K.; Shteamer, J.W.; Belyantseva, I.A.; Drummond, M.C.; Fitzgerald, T.S.; Vijayakumar, S.; Jones, S.M.; Griffith, A.J.; Friedman, T.B.; Cunningham, L.L.; et al. Gene Therapy Restores Balance and Auditory Functions in a Mouse Model of Usher Syndrome. Mol. Ther. Methods Clin. Dev. 2017, 25, 780–791. [Google Scholar] [CrossRef] [PubMed]
  73. Geng, R.; Omar, A.; Gopal, S.R.; Chen, D.H.C.; Stepanyan, R.; Basch, M.L.; Dinculescu, A.; Furness, D.N.; Saperstein, D.; Hauswirth, W.; et al. Modeling and Preventing Progressive Hearing Loss in Usher Syndrome III. Sci. Rep. 2017, 7, 13480. [Google Scholar] [CrossRef] [PubMed]
  74. Emptoz, A.; Michel, V.; Lelli, A.; Akil, O.; Boutet de Monvel, J.; Lahlou, G.; Meyer, A.; Dupont, T.; Nouaille, S.; Ey, E.; et al. Local gene therapy durably restores vestibular function in a mouse model of Usher syndrome type 1G. Proc. Natl. Acad. Sci. USA 2017, 114, 9695–9700. [Google Scholar] [CrossRef] [PubMed]
  75. Lee, M.Y.; Park, Y.-H. Potential of Gene and Cell Therapy for Inner Ear Hair Cells. Biomed. Res. Int. 2018, 2018, 8137614. [Google Scholar] [CrossRef] [PubMed]
  76. Petit, C.; Bonnet, C.; Safieddine, S. Deafness: From genetic architecture to gene therapy. Nat. Rev. Genet. 2023, 24, 665–686. [Google Scholar] [CrossRef] [PubMed]
  77. György, B.; Sage, C.; Indzhykulian, A.A.; Scheffer, D.I.; Brisson, A.R.; Tan, S.; Wu, X.; Volak, A.; Mu, D.; Tamvakologos, P.I.; et al. Rescue of Hearing by Gene Delivery to Inner-Ear Hair Cells Using Exosome-Associated AAV. Mol. Ther. 2017, 25, 379–391. [Google Scholar] [CrossRef] [PubMed]
  78. Akil, O.; Blits, B.; Lustig, L.R.; Leake, P.A. Virally Mediated Overexpression of Glial-Derived Neurotrophic Factor Elicits Age- and Dose-Dependent Neuronal Toxicity and Hearing Loss. Hum. Gene Ther. 2019, 30, 88–105. [Google Scholar] [CrossRef]
  79. French, L.S.; Mellough, C.B.; Chen, F.K.; Carvalho, L.S. A Review of Gene, Drug and Cell-Based Therapies for Usher Syndrome. Front. Cell Neurosci. 2020, 14, 183. [Google Scholar] [CrossRef]
  80. Delaney, D.S.; Liew, L.J.; Lye, J.; Atlas, M.D.; Wong, E.Y.M. Overcoming barriers: A review on innovations in drug delivery to the middle and inner ear. Front. Pharmacol. 2023, 14, 1207141. [Google Scholar] [CrossRef]
  81. Jinek, M.; Chylinski, K.; Fonfara, I.; Hauer, M.; Doudna, J.A.; Charpentier, E. A Programmable Dual-RNA–Guided DNA Endonuclease in Adaptive Bacterial Immunity. Science 2012, 337, 816–821. [Google Scholar] [CrossRef]
  82. Li, G.; Cheng, Y.; Li, Y.; Ma, H.; Pu, Z.; Li, S.; Zhao, Y.; Huang, X.; Yao, Y. A novel base editor SpRY-ABE8eF148A mediates efficient A-to-G base editing with a reduced off-target effect. Mol. Ther. Nucleic Acids 2023, 31, 78–87. [Google Scholar] [CrossRef] [PubMed]
  83. Walton, R.T.; Christie, K.A.; Whittaker, M.N.; Kleinstiver, B.P. Unconstrained genome targeting with near-PAMless engineered CRISPR-Cas9 variants. Science 2020, 368, 290–296. [Google Scholar] [CrossRef] [PubMed]
  84. Komor, A.C.; Badran, A.H.; Liu, D.R. CRISPR-Based Technologies for the Manipulation of Eukaryotic Genomes. Cell 2017, 168, 20–36. [Google Scholar] [CrossRef] [PubMed]
  85. Ran, F.A.; Hsu, P.D.; Wright, J.; Agarwala, V.; Scott, D.A.; Zhang, F. Genome engineering using the CRISPR-Cas9 system. Nat. Protoc. 2013, 8, 2281–2308. [Google Scholar] [CrossRef] [PubMed]
  86. Shams, F.; Bayat, H.; Mohammadian, O.; Mahboudi, S.; Vahidnezhad, H.; Soosanabadi, M.; Rahimpour, A. Advance trends in targeting homology-directed repair for accurate gene editing: An inclusive review of small molecules and modified CRISPR-Cas9 systems. BioImpacts BI 2022, 12, 371–391. [Google Scholar] [CrossRef] [PubMed]
  87. Gao, X.; Tao, Y.; Lamas, V.; Huang, M.; Yeh, W.-H.; Pan, B.; Hu, Y.-J.; Hu, J.H.; Thompson, D.B.; Shu, Y.; et al. Treatment of autosomal dominant hearing loss by in vivo delivery of genome editing agents. Nature 2018, 553, 217–221. [Google Scholar] [CrossRef]
  88. György, B.; Nist-Lund, C.; Pan, B.; Asai, Y.; Karavitaki, K.D.; Kleinstiver, B.P.; Garcia, S.P.; Zaborowski, M.P.; Solanes, P.; Spataro, S.; et al. Allele-specific gene editing prevents deafness in a model of dominant progressive hearing loss. Nat. Med. 2019, 25, 1123–1130. [Google Scholar] [CrossRef]
  89. Zheng, Z.; Li, G.; Cui, C.; Wang, F.; Wang, X.; Xu, Z.; Guo, H.; Chen, Y.; Tang, H.; Wang, D.; et al. Preventing autosomal-dominant hearing loss in Bth mice with CRISPR/CasRx-based RNA editing. Signal Transduct. Target. Ther. 2022, 7, 79. [Google Scholar] [CrossRef]
  90. Cui, C.; Wang, D.; Huang, B.; Wang, F.; Chen, Y.; Lv, J.; Zhang, L.; Han, L.; Liu, D.; Chen, Z.-Y.; et al. Precise detection of CRISPR-Cas9 editing in hair cells in the treatment of autosomal dominant hearing loss. Mol. Ther. Nucleic Acids 2022, 29, 400–412. [Google Scholar] [CrossRef]
  91. Noh, B.; Rim, J.H.; Gopalappa, R.; Lin, H.; Kim, K.M.; Kang, M.J.; Gee, H.Y.; Choi, J.Y.; Kim, H.H.; Jung, J. In vivo outer hair cell gene editing ameliorates progressive hearing loss in dominant-negative Kcnq4 murine model. Theranostics 2022, 12, 2465–2482. [Google Scholar] [CrossRef]
  92. Gu, X.; Hu, X.; Wang, D.; Xu, Z.; Wang, F.; Li, D.; Li, G.-l.; Yang, H.; Li, H.; Zuo, E.; et al. Treatment of autosomal recessive hearing loss via in vivo CRISPR/Cas9-mediated optimized homology-directed repair in mice. Cell Res. 2022, 32, 699–702. [Google Scholar] [CrossRef] [PubMed]
  93. Ingham, N.J.; Banafshe, N.; Panganiban, C.; Crunden, J.L.; Chen, J.; Lewis, M.A.; Steel, K.P. Inner hair cell dysfunction in Klhl18 mutant mice leads to low frequency progressive hearing loss. PLoS ONE 2021, 16, e0258158. [Google Scholar] [CrossRef]
  94. Fu, Y.; Foden, J.A.; Khayter, C.; Maeder, M.L.; Reyon, D.; Joung, J.K.; Sander, J.D. High-frequency off-target mutagenesis induced by CRISPR-Cas nucleases in human cells. Nat. Biotechnol. 2013, 31, 822–826. [Google Scholar] [CrossRef] [PubMed]
  95. Lin, Y.; Cradick, T.J.; Brown, M.T.; Deshmukh, H.; Ranjan, P.; Sarode, N.; Wile, B.M.; Vertino, P.M.; Stewart, F.J.; Bao, G. CRISPR/Cas9 systems have off-target activity with insertions or deletions between target DNA and guide RNA sequences. Nucleic Acids Res. 2014, 42, 7473–7485. [Google Scholar] [CrossRef] [PubMed]
  96. Gaudelli, N.M.; Komor, A.C.; Rees, H.A.; Packer, M.S.; Badran, A.H.; Bryson, D.I.; Liu, D.R. Programmable base editing of A•T to G•C in genomic DNA without DNA cleavage. Nature 2017, 551, 464–471. [Google Scholar] [CrossRef] [PubMed]
  97. Anzalone, A.V.; Randolph, P.B.; Davis, J.R.; Sousa, A.A.; Koblan, L.W.; Levy, J.M.; Chen, P.J.; Wilson, C.; Newby, G.A.; Raguram, A.; et al. Search-and-replace genome editing without double-strand breaks or donor DNA. Nature 2019, 576, 149–157. [Google Scholar] [CrossRef] [PubMed]
  98. Yeh, W.-H.; Chiang, H.; Rees, H.A.; Edge, A.S.B.; Liu, D.R. In vivo base editing of post-mitotic sensory cells. Nat. Commun. 2018, 9, 2184. [Google Scholar] [CrossRef]
  99. Yeh, W.H.; Shubina-Oleinik, O.; Levy, J.M.; Pan, B.; Newby, G.A.; Wornow, M.; Burt, R.; Chen, J.C.; Holt, J.R.; Liu, D.R. In vivo base editing restores sensory transduction and transiently improves auditory function in a mouse model of recessive deafness. Sci. Transl. Med. 2020, 12, 546. [Google Scholar] [CrossRef]
  100. Peters, C.W.; Hanlon, K.S.; Ivanchenko, M.V.; Zinn, E.; Linarte, E.F.; Li, Y.; Levy, J.M.; Liu, D.R.; Kleinstiver, B.P.; Indzhykulian, A.A.; et al. Rescue of hearing by adenine base editing in a humanized mouse model of Usher syndrome type 1F. Mol. Ther. 2023, 31, 2439–2453. [Google Scholar] [CrossRef]
  101. Tang, Z.-H.; Chen, J.-R.; Zheng, J.; Shi, H.-S.; Ding, J.; Qian, X.-D.; Zhang, C.; Chen, J.-L.; Wang, C.-C.; Li, L.; et al. Genetic Correction of Induced Pluripotent Stem Cells from a Deaf Patient with MYO7A Mutation Results in Morphologic and Functional Recovery of the Derived Hair Cell-Like Cells. Stem Cell Transl. Med. 2016, 5, 561–571. [Google Scholar] [CrossRef]
  102. Fuster-García, C.; García-García, G.; González-Romero, E.; Jaijo, T.; Sequedo, M.D.; Ayuso, C.; Vázquez-Manrique, R.P.; Millán, J.M.; Aller, E. USH2A Gene Editing Using the CRISPR System. Mol. Ther. Nucleic Acids 2017, 8, 529–541. [Google Scholar] [CrossRef]
  103. Zaw, K.; Wong, E.Y.M.; Zhang, X.; Zhang, D.; Chen, S.-C.; Thompson, J.A.; Lamey, T.; McLaren, T.; Roach, J.N.D.; Wilton, S.D.; et al. Generation of three induced pluripotent stem cell lines from a patient with Usher syndrome caused by biallelic c.949C>A and c.1256G>T mutations in the USH2A gene. Stem Cell Res. 2020, 50, 102129. [Google Scholar] [CrossRef] [PubMed]
  104. McLenachan, S.; Wong, E.Y.M.; Zhang, X.; Leith, F.; Moon, S.Y.; Zhang, D.; Chen, S.-C.; Thompson, J.A.; McLaren, T.; Lamey, T.; et al. Generation of two induced pluripotent stem cell lines from a patient with compound heterozygous mutations in the USH2A gene. Stem Cell Res. 2019, 36, 101420. [Google Scholar] [CrossRef] [PubMed]
  105. Sanjurjo-Soriano, C.; Erkilic, N.; Baux, D.; Mamaeva, D.; Hamel, C.P.; Meunier, I.; Roux, A.-F.; Kalatzis, V. Genome Editing in Patient iPSCs Corrects the Most Prevalent USH2A Mutations and Reveals Intriguing Mutant mRNA Expression Profiles. Mol. Ther. Methods Clin. Dev. 2020, 17, 156–173. [Google Scholar] [CrossRef] [PubMed]
  106. Meng, F.; Cang, X.; Peng, Y.; Li, R.; Zhang, Z.; Li, F.; Fan, Q.; Guan, A.S.; Fischel-Ghosian, N.; Zhao, X.; et al. Biochemical Evidence for a Nuclear Modifier Allele (A10S) in TRMU (Methylaminomethyl-2-thiouridylate-methyltransferase) Related to Mitochondrial tRNA Modification in the Phenotypic Manifestation of Deafness-associated 12S rRNA Mutation. J. Biol. Chem. 2017, 292, 2881–2892. [Google Scholar] [CrossRef] [PubMed]
  107. Chen, C.; Guan, M.-X. Genetic correction of TRMU allele restored the mitochondrial dysfunction-induced deficiencies in iPSCs-derived hair cells of hearing-impaired patients. Hum. Mol. Genet. 2022, 31, 3068–3082. [Google Scholar] [CrossRef] [PubMed]
  108. Farjami, M.; Assadi, R.; Afzal Javan, F.; Alimardani, M.; Eslami, S.; Mansoori Derakhshan, S.; Eslahi, A.; Mojarrad, M. The worldwide frequency of MYO15A gene mutations in patients with non-syndromic hearing loss: A meta-analysis. Iran. J. Basic. Med. Sci. 2020, 23, 841–848. [Google Scholar] [CrossRef]
  109. Chen, J.R.; Tang, Z.H.; Zheng, J.; Shi, H.S.; Ding, J.; Qian, X.D.; Zhang, C.; Chen, J.L.; Wang, C.C.; Li, L.; et al. Effects of genetic correction on the differentiation of hair cell-like cells from iPSCs with MYO15A mutation. Cell Death Differ. 2016, 23, 1347–1357. [Google Scholar] [CrossRef]
  110. Xue, Y.; Hu, X.; Wang, D.; Li, D.; Li, Y.; Wang, F.; Huang, M.; Gu, X.; Xu, Z.; Zhou, J.; et al. Gene editing in a Myo6 semi-dominant mouse model rescues auditory function. Mol. Ther. 2022, 30, 105–118. [Google Scholar] [CrossRef]
  111. Gu, X.; Wang, D.; Xu, Z.; Wang, J.; Guo, L.; Chai, R.; Li, G.; Shu, Y.; Li, H. Prevention of acquired sensorineural hearing loss in mice by in vivo Htra2 gene editing. Genome Biol. 2021, 22, 86. [Google Scholar] [CrossRef]
  112. Liu, X.; Lillywhite, J.; Zhu, W.; Huang, Z.; Clark, A.M.; Gosstola, N.; Maguire, C.T.; Dykxhoorn, D.; Chen, Z.-Y.; Yang, J. Generation and Genetic Correction of USH2A c.2299delG Mutation in Patient-Derived Induced Pluripotent Stem Cells. Genes 2021, 12, 805. [Google Scholar] [CrossRef] [PubMed]
  113. Kapustin, Y.; Chan, E.; Sarkar, R.; Wong, F.; Vorechovsky, I.; Winston, R.M.; Tatusova, T.; Dibb, N.J. Cryptic splice sites and split genes. Nucleic Acids Res. 2011, 39, 5837–5844. [Google Scholar] [CrossRef] [PubMed]
  114. Ham, K.A.; Keegan, N.P.; McIntosh, C.S.; Aung-Htut, M.T.; Zaw, K.; Greer, K.; Fletcher, S.; Wilton, S.D. Induction of cryptic pre-mRNA splice-switching by antisense oligonucleotides. Sci. Rep. 2021, 11, 15137. [Google Scholar] [CrossRef] [PubMed]
  115. Havens, M.A.; Hastings, M.L. Splice-switching antisense oligonucleotides as therapeutic drugs. Nucleic Acids Res. 2016, 44, 6549–6563. [Google Scholar] [CrossRef] [PubMed]
  116. Urban, E.; Noe, C.R. Structural modifications of antisense oligonucleotides. Farmaco 2003, 58, 243–258. [Google Scholar] [CrossRef] [PubMed]
  117. Shen, X.; Corey, D.R. Chemistry, mechanism and clinical status of antisense oligonucleotides and duplex RNAs. Nucleic Acids Res. 2018, 46, 1584–1600. [Google Scholar] [CrossRef]
  118. Summerton, J. Morpholino antisense oligomers: The case for an RNase H-independent structural type. Biochim. Biophys. Acta 1999, 1489, 141–158. [Google Scholar] [CrossRef]
  119. Grillet, N.; Xiong, W.; Reynolds, A.; Kazmierczak, P.; Sato, T.; Lillo, C.; Dumont, R.A.; Hintermann, E.; Sczaniecka, A.; Schwander, M.; et al. Harmonin mutations cause mechanotransduction defects in cochlear hair cells. Neuron 2009, 62, 375–387. [Google Scholar] [CrossRef]
  120. Tian, C.; Liu, X.Z.; Han, F.; Yu, H.; Longo-Guess, C.; Yang, B.; Lu, C.; Yan, D.; Zheng, Q.Y. Ush1c gene expression levels in the ear and eye suggest different roles for Ush1c in neurosensory organs in a new Ush1c knockout mouse. Brain Res. 2010, 1328, 57–70. [Google Scholar] [CrossRef]
  121. Lentz, J.; Pan, F.; Ng, S.S.; Deininger, P.; Keats, B. Ush1c216A knock-in mouse survives Katrina. Mutat. Res. 2007, 616, 139–144. [Google Scholar] [CrossRef]
  122. Lentz, J.J.; Jodelka, F.M.; Hinrich, A.J.; McCaffrey, K.E.; Farris, H.E.; Spalitta, M.J.; Bazan, N.G.; Duelli, D.M.; Rigo, F.; Hastings, M.L. Rescue of hearing and vestibular function by antisense oligonucleotides in a mouse model of human deafness. Nat. Med. 2013, 19, 345–350. [Google Scholar] [CrossRef] [PubMed]
  123. Boëda, B.; El-Amraoui, A.; Bahloul, A.; Goodyear, R.; Daviet, L.; Blanchard, S.; Perfettini, I.; Fath, K.R.; Shorte, S.; Reiners, J.; et al. Myosin VIIa, harmonin and cadherin 23, three Usher I gene products that cooperate to shape the sensory hair cell bundle. EMBO J. 2002, 21, 6689–6699. [Google Scholar] [CrossRef] [PubMed]
  124. Wang, L.; Kempton, J.B.; Jiang, H.; Jodelka, F.M.; Brigande, A.M.; Dumont, R.A.; Rigo, F.; Lentz, J.J.; Hastings, M.L.; Brigande, J.V. Fetal antisense oligonucleotide therapy for congenital deafness and vestibular dysfunction. Nucleic Acids Res. 2020, 48, 5065–5080. [Google Scholar] [CrossRef]
  125. Ponnath, A.; Depreux, F.F.; Jodelka, F.M.; Rigo, F.; Farris, H.E.; Hastings, M.L.; Lentz, J.J. Rescue of Outer Hair Cells with Antisense Oligonucleotides in Usher Mice Is Dependent on Age of Treatment. J. Assoc. Res. Otolaryngol. 2018, 19, 1–16. [Google Scholar] [CrossRef] [PubMed]
  126. Lentz, J.J.; Pan, B.; Ponnath, A.; Tran, C.M.; Nist-Lund, C.; Galvin, A.; Goldberg, H.; Robillard, K.N.; Jodelka, F.M.; Farris, H.E.; et al. Direct Delivery of Antisense Oligonucleotides to the Middle and Inner Ear Improves Hearing and Balance in Usher Mice. Mol. Ther. 2020, 28, 2662–2676. [Google Scholar] [CrossRef] [PubMed]
  127. Vaché, C.; Besnard, T.; le Berre, P.; García-García, G.; Baux, D.; Larrieu, L.; Abadie, C.; Blanchet, C.; Bolz, H.J.; Millan, J.; et al. Usher syndrome type 2 caused by activation of an USH2A pseudoexon: Implications for diagnosis and therapy. Hum. Mutat. 2012, 33, 104–108. [Google Scholar] [CrossRef]
  128. Slijkerman, R.W.N.; Vaché, C.; Dona, M.; García-García, G.; Claustres, M.; Hetterschijt, L.; Peters, T.A.; Hartel, B.P.; Pennings, R.J.E.; Millan, J.M.; et al. Antisense Oligonucleotide-based Splice Correction for USH2A-associated Retinal Degeneration Caused by a Frequent Deep-intronic Mutation. Mol. Ther. Nucleic Acids 2016, 5, e381. [Google Scholar] [CrossRef]
  129. Feng, P.; Xu, Z.; Chen, J.; Liu, M.; Zhao, Y.; Wang, D.; Han, L.; Wang, L.; Wan, B.; Xu, X.; et al. Rescue of mis-splicing of a common SLC26A4 mutant associated with sensorineural hearing loss by antisense oligonucleotides. Mol. Ther. Nucleic Acids 2022, 28, 280–292. [Google Scholar] [CrossRef]
  130. Panagiotopoulos, A.-L.; Karguth, N.; Pavlou, M.; Böhm, S.; Gasparoni, G.; Walter, J.; Graf, A.; Blum, H.; Biel, M.; Riedmayr, L.M.; et al. Antisense Oligonucleotide- and CRISPR-Cas9-Mediated Rescue of mRNA Splicing for a Deep Intronic CLRN1 Mutation. Mol. Ther. Nucleic Acids 2020, 21, 1050–1061. [Google Scholar] [CrossRef]
  131. Montgomery, M.K.; Xu, S.; Fire, A. RNA as a target of double-stranded RNA-mediated genetic interference in Caenorhabditis elegans. Proc. Natl. Acad. Sci. USA 1998, 95, 15502–15507. [Google Scholar] [CrossRef]
  132. Lam, J.K.; Chow, M.Y.; Zhang, Y.; Leung, S.W. siRNA Versus miRNA as Therapeutics for Gene Silencing. Mol. Ther. Nucleic Acids 2015, 4, e252. [Google Scholar] [CrossRef] [PubMed]
  133. Robillard, K.N.; de Vrieze, E.; van Wijk, E.; Lentz, J.J. Altering gene expression using antisense oligonucleotide therapy for hearing loss. Hear. Res. 2022, 426, 108523. [Google Scholar] [CrossRef] [PubMed]
  134. Kaur, T.; Mukherjea, D.; Sheehan, K.; Jajoo, S.; Rybak, L.P.; Ramkumar, V. Short interfering RNA against STAT1 attenuates cisplatin-induced ototoxicity in the rat by suppressing inflammation. Cell Death Dis. 2011, 2, e180. [Google Scholar] [CrossRef] [PubMed]
  135. Mukherjea, D.; Jajoo, S.; Kaur, T.; Sheehan, K.E.; Ramkumar, V.; Rybak, L.P. Transtympanic administration of short interfering (si)RNA for the NOX3 isoform of NADPH oxidase protects against cisplatin-induced hearing loss in the rat. Antioxid. Redox Signal 2010, 13, 589–598. [Google Scholar] [CrossRef] [PubMed]
  136. Oishi, N.; Chen, F.Q.; Zheng, H.W.; Sha, S.H. Intra-tympanic delivery of short interfering RNA into the adult mouse cochlea. Hear. Res. 2013, 296, 36–41. [Google Scholar] [CrossRef] [PubMed]
  137. Rodrigues, J.C.; Bachi, A.L.L.; Silva, G.A.V.; Rossi, M.; do Amaral, J.B.; Lezirovitz, K.; de Brito, R. New Insights on the Effect of TNF Alpha Blockade by Gene Silencing in Noise-Induced Hearing Loss. Int. J. Mol. Sci. 2020, 21, 2692. [Google Scholar] [CrossRef] [PubMed]
  138. Maeda, Y.; Fukushima, K.; Nishizaki, K.; Smith, R.J.H. In vitro and in vivo suppression of GJB2 expression by RNA interference. Hum Mol Genet. 2005, 14, 1641–1650. [Google Scholar] [CrossRef]
  139. De Vrieze, E.; Cañas Martín, J.; Peijnenborg, J.; Martens, A.; Oostrik, J.; van den Heuvel, S.; Neveling, K.; Pennings, R.; Kremer, H.; van Wijk, E. AON-based degradation of c.151C>T mutant COCH transcripts associated with dominantly inherited hearing impairment DFNA9. Mol. Ther. Nucleic Acids 2021, 24, 274–283. [Google Scholar] [CrossRef]
  140. Shibata, S.B.; Ranum, P.T.; Moteki, H.; Pan, B.; Goodwin, A.T.; Goodman, S.S.; Abbas, P.J.; Holt, J.R.; Smith, R.J.H. RNA Interference Prevents Autosomal-Dominant Hearing Loss. Am. J. Hum. Genet. 2016, 98, 1101–1113. [Google Scholar] [CrossRef]
  141. Yoshimura, H.; Shibata, S.B.; Ranum, P.T.; Moteki, H.; Smith, R.J.H. Targeted Allele Suppression Prevents Progressive Hearing Loss in the Mature Murine Model of Human TMC1 Deafness. Mol. Ther. 2019, 27, 681–690. [Google Scholar] [CrossRef]
  142. Du, X.; Li, W.; Gao, X.; West, M.B.; Saltzman, W.M.; Cheng, C.J.; Stewart, C.; Zheng, J.; Cheng, W.; Kopke, R.D. Regeneration of mammalian cochlear and vestibular hair cells through Hes1/Hes5 modulation with siRNA. Hear. Res. 2013, 304, 91–110. [Google Scholar] [CrossRef] [PubMed]
  143. Du, X.; Cai, Q.; West, M.B.; Youm, I.; Huang, X.; Li, W.; Cheng, W.; Nakmali, D.; Ewert, D.L.; Kopke, R.D. Regeneration of Cochlear Hair Cells and Hearing Recovery through Hes1 Modulation with siRNA Nanoparticles in Adult Guinea Pigs. Mol. Ther. 2018, 26, 1313–1326. [Google Scholar] [CrossRef] [PubMed]
  144. Zine, A.; Aubert, A.; Qiu, J.; Therianos, S.; Guillemot, F.; Kageyama, R.; de Ribaupierre, F. Hes1 and Hes5 Activities Are Required for the Normal Development of the Hair Cells in the Mammalian Inner Ear. J. Neurosci. 2001, 21, 4712–4720. [Google Scholar] [CrossRef] [PubMed]
  145. Takahashi, K.; Yamanaka, S. Induction of pluripotent stem cells from mouse embryonic and adult fibroblast cultures by defined factors. Cell 2006, 126, 663–676. [Google Scholar] [CrossRef] [PubMed]
  146. Takahashi, K.; Tanabe, K.; Ohnuki, M.; Narita, M.; Ichisaka, T.; Tomoda, K.; Yamanaka, S. Induction of Pluripotent Stem Cells from Adult Human Fibroblasts by Defined Factors. Cell 2007, 131, 861–872. [Google Scholar] [CrossRef] [PubMed]
  147. Shi, G.; Jin, Y. Role of Oct4 in maintaining and regaining stem cell pluripotency. Stem Cell Res. Ther. 2010, 1, 39. [Google Scholar] [CrossRef] [PubMed]
  148. Feng, R.; Wen, J. Overview of the roles of Sox2 in stem cell and development. Biol. Chem. 2015, 396, 883–891. [Google Scholar] [CrossRef] [PubMed]
  149. Nandan, M.O.; Yang, V.W. The role of Krüppel-like factors in the reprogramming of somatic cells to induced pluripotent stem cells. Histol. Histopathol. 2009, 24, 1343–1355. [Google Scholar] [CrossRef]
  150. Araki, R.; Hoki, Y.; Uda, M.; Nakamura, M.; Jincho, Y.; Tamura, C.; Sunayama, M.; Ando, S.; Sugiura, M.; Yoshida, M.A.; et al. Crucial Role of C-Myc in the Generation of Induced Pluripotent Stem Cells. Stem Cells 2011, 29, 1362–1370. [Google Scholar] [CrossRef]
  151. Li, H.; Roblin, G.; Liu, H.; Heller, S. Generation of hair cells by stepwise differentiation of embryonic stem cells. Proc. Natl. Acad. Sci. USA 2003, 100, 13495–13500. [Google Scholar] [CrossRef]
  152. Oshima, K.; Shin, K.; Diensthuber, M.; Peng, A.W.; Ricci, A.J.; Heller, S. Mechanosensitive Hair Cell-like Cells from Embryonic and Induced Pluripotent Stem Cells. Cell 2010, 141, 704–716. [Google Scholar] [CrossRef]
  153. Chen, W.; Jongkamonwiwat, N.; Abbas, L.; Eshtan, S.J.; Johnson, S.L.; Kuhn, S.; Milo, M.; Thurlow, J.K.; Andrews, P.W.; Marcotti, W.; et al. Restoration of auditory evoked responses by human ES-cell-derived otic progenitors. Nature 2012, 490, 278–282. [Google Scholar] [CrossRef]
  154. Barboza, L.C., Jr.; Lezirovitz, K.; Zanatta, D.B.; Strauss, B.E.; Mingroni-Netto, R.C.; Oiticica, J.; Haddad, L.A.; Bento, R.F. Transplantation and survival of mouse inner ear progenitor/stem cells in the organ of Corti after cochleostomy of hearing-impaired guinea pigs: Preliminary results. Braz. J. Med. Biol. Res. 2016, 49, e5064. [Google Scholar] [CrossRef]
  155. Chen, J.; Hong, F.; Zhang, C.; Li, L.; Wang, C.; Shi, H.; Fu, Y.; Wang, J. Differentiation and transplantation of human induced pluripotent stem cell-derived otic epithelial progenitors in mouse cochlea. Stem Cell Res. Ther. 2018, 9, 230. [Google Scholar] [CrossRef]
  156. Takeda, H.; Hosoya, M.; Fujioka, M.; Saegusa, C.; Saeki, T.; Miwa, T.; Okano, H.; Minoda, R. Engraftment of Human Pluripotent Stem Cell-derived Progenitors in the Inner Ear of Prenatal Mice. Sci. Rep. 2018, 8, 1941. [Google Scholar] [CrossRef]
  157. Lee, M.Y.; Hackelberg, S.; Green, K.L.; Lunghamer, K.G.; Kurioka, T.; Loomis, B.R.; Swiderski, D.L.; Duncan, R.K.; Raphael, Y. Survival of human embryonic stem cells implanted in the guinea pig auditory epithelium. Sci. Rep. 2017, 7, 46058. [Google Scholar] [CrossRef]
  158. Takeda, H.; Dondzillo, A.; Randall, J.A.; Gubbels, S.P. Selective ablation of cochlear hair cells promotes engraftment of human embryonic stem cell-derived progenitors in the mouse organ of Corti. Stem Cell Res. Ther. 2021, 12, 352. [Google Scholar] [CrossRef]
  159. Jeon, S.J.; Oshima, K.; Heller, S.; Edge, A.S. Bone marrow mesenchymal stem cells are progenitors in vitro for inner ear hair cells. Mol. Cell Neurosci. 2007, 34, 59–68. [Google Scholar] [CrossRef]
  160. Lee, J.H.; Kang, W.K.; Seo, J.H.; Choi, M.Y.; Lee, Y.H.; Kim, H.M.; Park, K.H. Neural differentiation of bone marrow-derived mesenchymal stem cells: Applicability for inner ear therapy. Korean J. Audiol. 2012, 16, 47–53. [Google Scholar] [CrossRef]
  161. Jang, S.; Cho, H.H.; Kim, S.H.; Lee, K.H.; Jun, J.Y.; Park, J.S.; Jeong, H.S.; Cho, Y.B. Neural-induced human mesenchymal stem cells promote cochlear cell regeneration in deaf Guinea pigs. Clin. Exp. Otorhinolaryngol. 2015, 8, 83–91. [Google Scholar] [CrossRef]
  162. Mittal, R.; Ocak, E.; Zhu, A.; Perdomo, M.M.; Pena, S.A.; Mittal, J.; Bohorquez, J.; Eshraghi, A.A. Effect of Bone Marrow-Derived Mesenchymal Stem Cells on Cochlear Function in an Experimental Rat Model. Anat. Rec. 2020, 303, 487–493. [Google Scholar] [CrossRef]
  163. Tsai, S.C.; Yang, K.D.; Chang, K.H.; Lin, F.C.; Chou, R.H.; Li, M.C.; Cheng, C.C.; Kao, C.Y.; Chen, C.P.; Lin, H.C.; et al. Umbilical Cord Mesenchymal Stromal Cell-Derived Exosomes Rescue the Loss of Outer Hair Cells and Repair Cochlear Damage in Cisplatin-Injected Mice. Int. J. Mol. Sci. 2021, 22, 6664. [Google Scholar] [CrossRef]
  164. Huang, X.; Kou, X.; Zhan, T.; Wei, G.; He, F.; Mao, X.; Yang, H. Apoptotic vesicles resist oxidative damage in noise-induced hearing loss through activation of FOXO3a-SOD2 pathway. Stem Cell Res. Ther. 2023, 14, 88. [Google Scholar] [CrossRef]
  165. Warnecke, A.; Prenzler, N.; Harre, J.; Köhl, U.; Gärtner, L.; Lenarz, T.; Laner-Plamberger, S.; Wietzorrek, G.; Staecker, H.; Lassacher, T.; et al. First-in-human intracochlear application of human stromal cell-derived extracellular vesicles. J. Extracell. Vesicles 2021, 10, e12094. [Google Scholar] [CrossRef]
  166. Galán, L.; Guerrero-Sola, A.; Gómez-Pinedo, U.; Matias-Guiu, J. Cell therapy in amyotrophic lateral sclerosis: Science and controversy. Neurología 2010, 25, 467–469. [Google Scholar] [CrossRef]
  167. Baumgartner, L.S.; Moore, E.; Shook, D.; Messina, S.; Day, M.C.; Green, J.; Nandy, R.; Seidman, M.; Baumgartner, J.E. Safety of Autologous Umbilical Cord Blood Therapy for Acquired Sensorineural Hearing Loss in Children. J. Audiol. Otol. 2018, 22, 209–222. [Google Scholar] [CrossRef]
  168. Lee, H.S.; Kim, W.J.; Gong, J.S.; Park, K.H. Clinical Safety and Efficacy of Autologous Bone Marrow-Derived Mesenchymal Stem Cell Transplantation in Sensorineural Hearing Loss Patients. J. Audiol. Otol. 2018, 22, 105–109. [Google Scholar] [CrossRef]
  169. Andrade da Silva, L.H.; Heuer, R.A.; Roque, C.B.; McGuire, T.L.; Hosoya, T.; Kimura, H.; Tamura, K.; Matsuoka, A.J. Enhanced survival of hypoimmunogenic otic progenitors following intracochlear xenotransplantation: Repercussions for stem cell therapy in hearing loss models. Stem Cell Res. Ther. 2023, 14, 83. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the sensory HCs and auditory nerves of the cochlea. In the normal organ of Corti, the HCs have upright stereocilia and are innervated with auditory neurons surrounding their base. Various insults (e.g., noise, ototoxic drugs and/or genetic) can damage or cause cell death of the sensory cells resulting in SNHL. Current therapeutic strategies target the signaling pathways and/or alter the gene expression of the affected cells to restore these cells. These approaches can (i) reprogram the cell fate of SCs into cochlear HCs by transdifferentiation, (ii) correct or disrupt the pathogenic gene by genome editing (e.g., CRISPR-Cas systems), modulate gene expression (e.g., ASO), or deliver normal gene transcripts via viral vectors (e.g., AAV), and (iii) use cell-based therapy in combination with gene targeted therapy to replace the damaged HCs and nerves.
Figure 1. Schematic representation of the sensory HCs and auditory nerves of the cochlea. In the normal organ of Corti, the HCs have upright stereocilia and are innervated with auditory neurons surrounding their base. Various insults (e.g., noise, ototoxic drugs and/or genetic) can damage or cause cell death of the sensory cells resulting in SNHL. Current therapeutic strategies target the signaling pathways and/or alter the gene expression of the affected cells to restore these cells. These approaches can (i) reprogram the cell fate of SCs into cochlear HCs by transdifferentiation, (ii) correct or disrupt the pathogenic gene by genome editing (e.g., CRISPR-Cas systems), modulate gene expression (e.g., ASO), or deliver normal gene transcripts via viral vectors (e.g., AAV), and (iii) use cell-based therapy in combination with gene targeted therapy to replace the damaged HCs and nerves.
Biomedicines 11 03347 g001
Table 1. Current clinical trials of drug interventions for inner ear hearing disorders.
Table 1. Current clinical trials of drug interventions for inner ear hearing disorders.
Cause of Hearing LossIntervention (Drug Name)Route of AdministrationSponsor (Company)Drug Component (Mechanism of Action)Study Phase and Status (Clinical Trial ID)
ARHLSmall molecule
(PF-04958242)
Oral solutionBiogen (Cambridge, MA, USA)AMPA receptor potentiatorPhase 1 completed (NCT01518920)
ARHL, tinnitusSmall molecule
(AUT00063)
Oral administrationAutifony Therapeutics Limited (Stevenage, UK)Kv3 potassium channel modulatorPhase 2 completed (NCT02315508; NCT02345031; NCT02832128)
AIEDBiologics
(Anakinra (Kineret))
Subcutaneous injectionNorthwell Health (New Hyde Park, NY, USA)Interleukin-1 receptor antagonistPhase 2 (NCT03587701)
AIEDMonoclonal antibody
(Gevokizumab)
Subcutaneous injectionXOMA Corporation (Emeryville, CA, USA)Proinflammatory interleukin-1 inhibitorPhase 2 completed (NCT01950312)
AIEDFusion protein
(Rilonacept)
Subcutaneous injectionStanley Cohen (Regeneron Pharmaceuticals) (Tarrytown, NY, USA)Interleukin-1 receptor antagonistPhase 1 (NCT02828033)
DIHL (aminoglycoside)Small molecule
(ORC-13661)
Oral administrationKevin Winthrop (Oricula Therapeutics) (Seattle, WA, USA)Mechanotransduction channel blockerPhase 2 (NCT05730283)
DIHL (cisplatin)Small molecule
(DB-020)
IntratympanicDecibel Therapeutics (Boston, MA, USA)Sodium thiosulfate—cation chelator and antioxidant Phase 1 completed (NCT04262336)
DIHL (cisplatin)Small moleculeIntravenousHyunseok Kang (NRG Oncology) (Philadelphia, PA, USA)Sodium thiosulfate—cation chelator and antioxidantPhase 2 completed (NCT04541355)
DIHL (cisplatin)Small molecule Intravenous infusionSunnybrook Health Sciences Centre (Toronto, ON, Canada)Sodium thiosulfate—cation chelator and antioxidant; Mannitol—JNK pathway inhibitor Phase 2 (NCT05129748)
DIHL (cisplatin)Small molecule IntratympanicSunnybrook Health Sciences Centre (Toronto, ON, Canada)N-acetylcysteine—antioxidant, reduce inflammationPhase 1/2 (NCT04291209)
DIHLSmall molecule
(SPI-3005)
Oral administrationSound Pharmaceuticals (Seattle, WA, USA) Ebselen and allopurinol—mimic antioxidant property of glutathione peroxidasePhase 2 (NCT01451853; NCT02819856)
Congenital CMV HLAntiviral
(Valganciclovir)
Oral solutionDr. Ann C.T.M. Vossen (Stichting Nuts Ohra (Amsterdam, The Netherlands) and Leiden University Medical Center (Leiden, The Netherlands))Viral DNA synthesis inhibitorPhase 3 terminated (NCT01655212); Phase 3 completed (NCT02005822)
Congenital CMV HLAntiviral
(Valganciclovir)
Oral solutionAlbert Park (Genentech) (South San Francisco, CA, USA)Viral DNA synthesis inhibitorPhase 2 (NCT03107871)
Congenital CMV HLmRNA vaccine
(mRNA-1647)
Intramuscular injection ModernaTX, Inc. (Cambridge, MA, USA) Prophylactic vaccine containing six mRNA sequences for two CMV antigens (glycoprotein B and pentameric glycoprotein complex)Phase 1 completed (NCT05105048); Phase 2 (NCT0683457); Phase 3 (NCT05085366)
MD, NIHL, tinnitusSmall molecule
(SPI-1005)
Oral administrationSound Pharmaceuticals (Seattle, WA, USA)Ebselen—mimic antioxidant property of glutathione peroxidasePhase 1 completed (NCT01452607); Phase 1/2 completed (NCT02603081); Phase 2 (NCT02779192); Phase 2 completed (NCT01444846; NCT03325790); Phase 3 (NCT04677972)
MDSmall molecule
(Latanoprost)
IntratympanicSynphora AB (Uppsala, Sweden) Prostaglandin F2 α receptor agonistPhase 2 (NCT01973114)
MDSmall molecule
(Montelukast)
Oral administrationHouse Ear Institute (Merck) (Los Angeles, CA, USA)Leukotriene receptor antagonists Phase 4 (NCT04815187)
NIHLSmall molecule
(Zonisamide)
Oral administrationGateway Biotechnology (St. Louis, MO, USA)Voltage-dependent sodium channel inhibitor and neurotransmitters degrader Phase 2 (NCT04768569); Phase 2 terminated (NCT04774250)
NIHL, SNHLSmall molecule
(NHPN-1010)
Oral administrationOtologic Pharmaceuticals (Oklahoma City, OK, USA)Disufenton—nitrone-based antioxidant and neuroprotective; N-acetylcysteine—free radical scavenger Phase 1 completed (NCT02259595)
NIHL (with mitochondrial point mutation)Small molecule
(VincerinoneTM (EPI-743))
Oral administrationEdison Pharmaceuticals (Mountain View, CA, USA)Vatiquinone—mitochondrial redox modulator Phase 2 completed (NCT02257983)
SNHLSmall molecule
(AM-111)
IntratympanicAuris Medical (Basel, Switzerland)Brimapitide—JNK pathway inhibitorPhase 2 completed (NCT00802425); Phase 3 completed (NCT02561091); Phase 3 terminated (NCT02809118)
SNHLSmall molecule
(PIPE-505)
IntratympanicPipeline Therapeutics (San Diego, CA, USA)HC regeneration via Notch signaling pathway and γ-secretase inhibition; synaptic regeneration via netrin/DCC pathway inhibitionPhase 1/2 completed (NCT04462198)
SNHLCell-based therapy
(Autologous bone marrow infusion)
Intravenous infusionJames Baumgartner, CBR Systems, Inc. (Los Angeles, CA, USA)Umbilical cord-derived mesenchymal stem cells Phase 1/2 completed (NCT02038972)
Idiopathic SSNHLSteroid
(AC102)
Intratympanic AudioCure Pharma GmbH (Berlin, Germany)Prednisolone Phase 2 (NCT05776459)
SSNHLSteroid IntratympanicMassachusetts Eye and Ear Infirmary (Boston, MA, USA)Dexamethasone sodium succinate and prednisone Phase 3 completed (NCT00097448)
SSNHLSteroidIntratympanicUniversity Hospital Tuebingen (Tuebingen, Germany)Dexamethasone-dihydrogenphosphatePhase 3 completed (NCT00335920)
SSNHLBiologics
(Ancrod)
IntravenousNordmark Arzneimittel GmbH & Co. KG (Uetersen, Germany)Anticoagulant—fibrinogen inhibitors and plasminogen activator stimulantsPhase 1/2 completed (NCT01621256)
SSNHLSteroid Oral administrationUniversity of Colorado, Denver (Denver, CO, USA)Dexamethasone and prednisone Phase 2 completed (NCT03255473)
SSNHLSteroidOral administrationBeijing Tsinghua Chang Gung Hospital (Beijing, USA)Methylprednisolone hemisuccinate and ginkgo biloba Phase 4 (NCT04192656)
SSNHL SteroidIntratympanic with hydrogel gel Seoul National University Hospital (Seoul, South Korea)Dexamethasone and hyaluronic acid Phase 1/2 (NCT04766853)
SSNHLSteroidOral administrationNorthwestern University (Evanston, IL, USA)Dexamethasone and methylprednisolone Phase 4 (NCT04826237)
SSNHLSmall molecule
(STR001)
Intratympanic versus oral administrationStrekin AG (Basel, Switzerland)Peroxisome proliferator-activated receptor-γ agonistPhase 3 completed (NCT03331627)
SSNHLSmall molecule
(SENS-401)
Oral administrationSensorion (Montpellier, France)R-azasetron besylate—5-HT3 antagonist and calcineurin inhibitorPhase 2 (NCT05258773; NCT05628233); Phase 2/3 completed (NCT03603314)
SSNHL Steroid
(HY01)
IntratympanicHeyu (Suzhou, China) Pharmaceutical Technology Co. (Wuxi, China)Dexamethasone sodium phosphate Phase 1 (NCT04961099)
TinnitusSmall molecule
(Brexanolone)
IntravenousSage Therapeutics (Cambridge, MA, USA)Gamma-aminobutyric acid A receptor positive modulatorPhase 2 (NCT05645432)
TinnitusSmall molecule
(Gabapentin)
Oral administrationIslamic Azad University of Mashhad (Mashhad, Iran)Gamma-aminobutyric acid analogue Phase 2 completed (NCT00555776)
Tinnitus Small molecule
(BGG492A (Selurampanel))
Oral administrationNovartis Pharmaceuticals (Basel, Switzerland)AMPA-type glutamate receptor antagonist Phase 2 completed (NCT01302873)
TinnitusSmall molecule
(Etanercept–Enbrel)
Systemic injectionWayne State University (Detroit, MI, USA)TNF-α inhibitorPhase 2 (NCT04066348)
Genetic (otoferlin-mediated HL)Gene therapy
(AK-OTOF)
IntracochlearAkouos, Inc. (Boston, MA, USA) Replace defective OTOF with healthy OTOF cDNA transcriptPhase 1/2 (NCT05821959)
Genetic (otoferlin-mediated HL)Gene therapy
(DB-OTO)
IntracochlearDecibel Therapeutics (Boston, MA, USA)Replace defective OTOF with healthy OTOF cDNA transcriptPhase 1/2 (NCT05788536)
Genetic (otoferlin-mediated HL)Gene editing
(HG205)
IntracochlearHuidaGene Therapeutics Co., Ltd. (Shangai, China)CRISPR/Cas13 RNA base editing for p.Q829X mutation in the OTOF genePhase 1 (NCT06025032)
Genetic (Wolfram syndrome)Small molecule
(Depakine)
Oral administrationCentre d’Etude des Cellules Souches (Corbeil-Essonnes, France)Sodium valproate—voltage-gated ion channel blocker and histone deacetylase inhibitorPhase 2 (NCT04940572)
Genetic (mitochondrial DNA tRNA mutation) Small molecule
(KH176)
Oral administration Khondrion BV (Nijmegen, The Netherlands)Intracellular redox modulating agent Phase 2 (NCT04604548)
Note: The database search was conducted as of 18 October 2023, using the key terms (hearing disorder), (sensorineural hearing loss) and (inner ear). Abbreviations: AIED, autoimmune inner ear disease; ARHL, age-related hearing loss; CMV HL, cytomegalovirus-infection-induced hearing loss; DIHL, drug-induced hearing loss; HC, hair cell; MD, Ménière’s disease; NIHL, noise-induced hearing loss; SSNHL, sudden SNHL.
Table 2. Preclinical therapeutic pipeline for inner ear hearing disorders.
Table 2. Preclinical therapeutic pipeline for inner ear hearing disorders.
Cause of Hearing LossTherapeutic EffectIntervention (Drug Name)Route of AdministrationSponsor (Company)Drug Component (Mechanism of Action)
ARHL, NIHL OtoprotectionACOU082Transdermal patchAcousia Therapeutics (Tuebingen, Germany)KCNQ4 receptor agonist
DIHL (cisplatin)OtoprotectionACOU085IntratympanicAcousia Therapeutics (Tuebingen, Germany)KCNQ4 receptor agonist
Genetic (CLRN1-mediated HL)Gene therapyAK-CLRN1IntratympanicAkouos, Inc. (Boston, MA, USA)AAV CLRN1 gene transfer to cochlear HCs
Genetic (autosomal dominant)Gene therapyUndisclosedN/AAkouos, Inc. (Boston, MA, USA)N/A
Genetic (GJB2-mediated HL)Gene therapyGJB2N/AAkouos, Inc. (Boston, MA, USA)GJB2 gene correction in SCs
Genetic HC regenerationUndisclosedN/AAkouos, Inc. (Boston, MA, USA)N/A
Genetic (vestibular schwannoma)Gene therapyAK-antiVEGFN/AAkouos, Inc. (Boston, MA, USA)N/A
AIHL, DIHL, NIHL OtoprotectionAP-001Transtympanic Anida Pharma (Cambridge, MA, USA) Neuroprotectin (D1/NPD1) with anti-inflammatory, cell survival and tissue repair properties
Genetic (GJB2-mediated HL)Gene therapyAAV.103N/ADecibel Therapeutics (collaboration with Regeneron)AAV GJB2 gene transfer to cells that would normally express GJB2
Genetic (STRC-mediated HL)Gene therapyAAV.104UndisclosedDecibel Therapeutics (collaboration with Regeneron)AAV STRC gene transfer to selectively target outer HCs
GeneticGene therapyAAV.105UndisclosedDecibel Therapeutics (Boston, MA, USA)Undisclosed
GeneticRegeneration AAV.201UndisclosedDecibel Therapeutics (Boston, MA, USA)ATOH1 and SOX2 inhibitor
GeneticHC regenerationUndisclosedUndisclosedDecibel Therapeutics (Boston, MA, USA)Undisclosed
TinnitusUndisclosedGW-TT2Nasal sprayGateway Biotechnology (St. Louis, MO, USA)FDA-approved drug
TinnitusGene therapy GW-TT5N/AGateway Biotechnology (St Louis, MO, USA)N/A
TinnitusUndisclosedGW-TT23Nasal sprayGateway Biotechnology (St. Louis, MO, USA)N/A
DIHL, NIHLNeural protection and regeneration HB-097
(Cometin)
N/AHoba Therapeutics (Copenhagen, Denmark)Neurotrophic factor targeting the JAK-STAT3 and MEK-MERK pathways promote the survival of neuronal cells
MDOtoprotectionIB2000N/AIntraBio (London, UK)Diferuloylmethane analogue with anti-inflammatory property
MDOtoprotectionIB5000
(Betahistidine)
N/AIntraBio (London, UK)Monoamine oxidase inhibitor
Auditory neuropathy Neural cell replacement therapy ANP1N/ALineage Cell Therapeutics (Carlsbad, CA, USA)Replace damaged cells with auditory neuronal progenitors
Genetic (TMPRSS3-mediated HL)Gene therapy Myr-201N/AMyrtelle, Inc. (Wakefield, MA, USA)AAV TMPRSS3 gene transfer
AIEDOtoprotectionOR-102AN/AO-Ray Pharma (Pasadena, CA, USA)TNFα inhibitor
SNHLOtoprotectionOR-102CN/AO-Ray Pharma (Pasadena, CA, USA)Protective effect during cochlear implant surgery
ARHLOtoprotection OR-112N/AO-Ray Pharma (Pasadena, CA, UAS)N/A
MDOtoprotectionORB-202
(Betamethasone)
Intratympanic Orbis Biosciences (Lenexa, KS, USA)Fast film-forming agent (FFA) encapsulating betamethasone
SNHL HC regeneration Small molecule
(OPI-001)
N/AOtologic Pharmaceuticals (Okhaloma City, OK, USA)Small molecule drugs and siRNA to promote HC regeneration
Genetic (TMPRSS3-mediated HL)Gene therapyRHI100N/ARescue Hearing Inc. (Gainesville, FL, USA)AAV TMPRSS3 gene transfer
Genetic (non-syndromic HL)Gene therapyRHI400N/ARescue Hearing Inc. (Gainesville, FL, USA)Target major cause of non-syndromic HL
SNHLGene therapyRHI500N/ARescue Hearing Inc. (Gainesville, FL, USA)Target HL in cochlear implant population
Neurodegenerative disease Gene therapyRHI600N/ARescue Hearing Inc (Gainesville, FL, USA)N/A
Auditory neuropathy Neural cell replacement therapy Rincell-1N/ARinri Therapeutics (Sheffield, UK)Replace damaged cells with embryonic stem cells
Genetic (OTOF-mediated HL)Gene therapyOTOF-GT
(SENS-501)
N/ASensorion (Montpellier, France)Dual AAV OTOF gene transfer to cochlear HCs
Genetic (GJB2-mediated HL)Gene therapyGJB2-GTN/ASensorion (Montpellier, France)Dual AAV GJB2 gene transfer to cochlear HCs
SNHLRegenerationSPI-5557N/ASound Pharmaceuticals (Seattle, WA, USA)Cyclin-dependent kinase (p27Kip1) inhibitor
DIHL (cisplatin)OtoprotectionLPT99Transtympanic (hydrogel formulation)Spiral Therapeutics (South San Francisco, CA, USA)APAF-1 inhibitor, anti-apoptotic
DIHL (cisplatin)OtoprotectionTT001
(AZD5438)
N/ATing Therapeutics (Ohama, NE, USA)Cyclin-dependent kinase (p27Kip1) inhibitor
DIHL (cisplatin)Otoprotection TT002
(Niclosamide)
N/ATing Therapeutics (Ohama, NE, USA)EGFR-ERK pathway
DIHL Otoprotection TT003
(Piperlongumine)
N/ATing Therapeutics (Ohama, NE, USA)Akt phophorylation inhibition via the accumulation of reactive oxygen species
Abbreviations: AIED, autoimmune inner ear disease; ARHL, age-related hearing loss; DIHL, drug-induced hearing loss; HC; hair cell; HL, hearing loss; MD, Ménière’s disease; NIHL, noise-induced hearing loss; SC, supporting cell; SSNHL, sudden SNHL.
Table 4. Antisense oligonucleotide-based therapy for SNHL in preclinical models.
Table 4. Antisense oligonucleotide-based therapy for SNHL in preclinical models.
Gene/Disease ModelModel Mechanism of ActionChemical ModificationEffectReference
DFNA9Minigene splicing assay RNAse H1-mediated degradation with siRNA 2′-deoxyribose flanked with 2′-O-methyl-RNA with PS backbone Selective suppression of the mutant COCH c.151C>T pre-mRNA transcript[139]
DFNA3Gjb2R75W mice, P19 RNAi-mediated degradation with siRNA 2′-deoxythymidine residues Selective suppression of the dominant negative Gjb2R75W expression in the organ of Corti and prevention of HL in mice [140]
Pendred syndromePatient-derived PBM cells; chimeric mice, P3–P4Splice modification 2′-MOE with PS backboneCorrection of splicing caused by SLC26A4 c.919-2A>G mutation in patient-derived blood cells and in the chimeric mouse inner ear [129]
DFNA36Beethoven mice, P0–P2RNAi-mediated degradation with miRNAmiRNA conjugated to AAV2/9 vectorImproved ABR thresholds; expression of Tmc1Bth/+ mutant allele significantly reduced compared to the untreated mutant Bth mice [140]
DFNA36Beethoven mice, P15–P16, P56–P60 and P84–P90RNAi-mediated degradation with miRNA miRNA conjugated to AAV2/9 vector Single dose of miTmc at P15–P16 showed significant preservation of hearing and HC morphology at 8–12 weeks of age [141]
Usher syndrome type 1CUsh1c216AA mice, P3–P18Splice modification 2′-MOE with PS backbone Improved hearing function and correction of harmonin expression in HCs was maintained up to P180 in Ush1c mutant mice injected at P3 and P5 [122]
Usher syndrome type 1CUsh1c216AA mice, P1–P7Splice modification 2′-MOE with PS backbone ASO-29 treatment at P1–P5 rescued both IHCs and OHCs but treatment at P7 only partially rescued IHC function[125]
Usher syndrome type 1CUsh1c216AA mice, P1–P20 Splice modification 2′-MOE with PS backbone Detectable harmonin protein expression and correction of full-length harmonin transcript;
post-treatment of ASO-29 at 1 month showed significant improvement in hearing
[126]
Usher syndrome type 1CUsh1c216AA mice, E12.5 Splice modification 2′-MOE with PS backbone Partial restoration of hearing and vestibular function was observed 1 month after injecting in utero in E12.5 Ush1c mice [124]
Usher syndrome type 2APatient-derived fibroblasts (USH2A c.7595-2144A>G and c.12575G>A) Splice modification 2′-MOE with PS backbone ASO treatment excluded intron 40 caused by the USH2A c.7595-2144A>G mutation into the mature USH2A mRNA transcript in patient-derived fibroblasts[128]
Usher syndrome type 3AMinigene splicing assay Splice modification 2′-MOE with PS backbone ASO treatment showed robust splicing correction of the CLRN1 c.254-649T>G splicing mutation in the cell lines carrying patient mutation[130]
Cisplatin-induced HLCisplatin-treated ratsRNA-mediated degradation with siRNAN/ANOX3 siRNA treatment showed otoprotective effects by suppressing inflammation and apoptotic activity in the cochlea[135]
Cisplatin-induced HLCisplatin-treated ratsRNAi-mediated degradation with siRNAN/A Knockdown of STAT1 inhibited cisplatin-induced inflammation and protected HCs in rat cochleae[134]
NIHLNoise-exposed miceRNAi-mediated degradation with siRNAN/ANOX3 expression was reduced in OHCs and hearing threshold shifts were attenuated in noise-exposed mice[136]
NIHLNoise-exposed rats RNAi-mediated degradation with siRNAN/A Knockdown of TNFα downregulated the genes involved in ROS generation, and therefore improved the hearing outcome after noise exposure [137]
HC regenerationNeomycin-treated mouse cochlea explants RNAi-mediated degradation with siRNAN/AHes1 and Hes5 mRNA expression levels were reduced where an increase in Atoh1 expression level led to HC regeneration [142]
HC regenerationNoise-exposed guinea pigs RNAi-mediated degradation with siRNAN/ASupernumerary IHCs were observed in 50% of the cochleae treated with Hes1 siRNA; ABR thresholds were well preserved [143]
2-MOE, 2′-O-methoxyethyl; HC, hair cell; HL, hearing loss; IHC, inner hair cell; NIHL, noise-induced hearing loss; OHC, outer hair cell; PBM, peripheral blood mononuclear, PS, phosphorothioate; ROS, reactive oxygen species.
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

Lye, J.; Delaney, D.S.; Leith, F.K.; Sardesai, V.S.; McLenachan, S.; Chen, F.K.; Atlas, M.D.; Wong, E.Y.M. Recent Therapeutic Progress and Future Perspectives for the Treatment of Hearing Loss. Biomedicines 2023, 11, 3347. https://doi.org/10.3390/biomedicines11123347

AMA Style

Lye J, Delaney DS, Leith FK, Sardesai VS, McLenachan S, Chen FK, Atlas MD, Wong EYM. Recent Therapeutic Progress and Future Perspectives for the Treatment of Hearing Loss. Biomedicines. 2023; 11(12):3347. https://doi.org/10.3390/biomedicines11123347

Chicago/Turabian Style

Lye, Joey, Derek S. Delaney, Fiona K. Leith, Varda S. Sardesai, Samuel McLenachan, Fred K. Chen, Marcus D. Atlas, and Elaine Y. M. Wong. 2023. "Recent Therapeutic Progress and Future Perspectives for the Treatment of Hearing Loss" Biomedicines 11, no. 12: 3347. https://doi.org/10.3390/biomedicines11123347

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