uc sf · auditory hair cell replacement and hearing improvement by atoh1 gene therapy in deaf...

30
11/8/2013 1 Lawrence R. Lustig, MD Department of Oto-HNS University of California San Francisco Cochlear Gene Therapy: Is It Time? UC SF > 1700 Gene Therapy Trials Retinal Congenital Amaurosis X-linked SCID ADA-SCID Adrenoleukodystrophy Parkinson’s Disease HIV Is It Time for the Ear?

Upload: others

Post on 27-Feb-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

1

Lawrence R. Lustig, MDDepartment of Oto-HNS

University of California San Francisco

Cochlear Gene Therapy: Is It Time?

UCSF

> 1700 Gene Therapy Trials

• Retinal Congenital Amaurosis• X-linked SCID• ADA-SCID• Adrenoleukodystrophy• Parkinson’s Disease• HIV Is It Time for the Ear?

Page 2: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

2

Vector

Promoter

Gene

Which Vector?

Vector Advantages Disadvantages

Adenovirus Easy to make<10kb inserts

ImmunogenicTransient expression

Adeno-Assoc Virus(AAV)

Non-pathologicCell division not req’d

<5kb insertsDifficult to make

Variable transfection

Herpesvirus 100kb insertsStable expression

Human diseasecytopathic

Lentivirus Stable expression Insertional mutagenesisLow transfection rate

Liposomes Easy to makeNo insert size limit

Non-pathologic

Low and transient expression

Which Vector?

Vector Advantages Disadvantages

Adenovirus Easy to make<10kb inserts

ImmunogenicTransient expression

Adeno-Assoc Virus Non-pathologicCell division not req’d

<5kb insertsDifficult to make

Variable transfection

Herpesvirus 100kb insertsStable expression

Human diseasecytopathic

Lentivirus Stable expression Insertional mutagenesisLow transfection rate

Liposomes Easy to makeNo insert size limit

Non-pathologic

Low and transient expression

Which Promoter?

Promoter

Gene

• “Constitutive” = Always on

• Gene-specific / Cell Specific

Page 3: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

3

Which Gene ???

Promoter

Gene

• Hair Cell Regeneration?

• Neuronal Growth Factors?

• Antioxidant / Antiapoptic factors?

• AAV-β-gal, Ad2 major late promoter

• Osmotic minipump

• Signal in spiral limbus, spiral ligament, spiral ganglion cells and the organ of Corti

• Weaker, similar pattern in contra ear

Development of in vivo gene therapy for hearing disorders: introduction of adeno-associated virus into the cochlea of the guinea pig.

Lalwani et al., Gene Therapy, 1996

•AAV-GFP

•CMV (constitutive) Promoter

•5 Serotypes studied (1, 2, 5, 6, 8)

•+/- deafeningKilpatrick et al, 2011

•Transduced:Hair cells, supporting cellsauditory nervespiral ligament

•IHCs most effectively transduced

•All 5 serotypes transducedauditory nerve

•AAV-8 most efficient vector

Page 4: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

4

2. RoundWindow Injection

Routes of Delivery

3. Cochleostomy

1. RoundWindow Diffusion

RWM Diffusion

•Variable carrier (ie gelfoam)

•Poor uptake alone

•Improved w/ Hyaluronic AcidShibata et al, Hum Gene Therapy 2011

•Enzymatic Digestion of RWWang et al, Gene Therapy, 2011

RWM Injection vs CochleostomyRWM Injection Cochleostomy

Less traumaticEasier approach

Theoretic access to SM

Direct access to SM

Used in human CI Used in human CI

Cochlear Gene Therapy Applications

Page 5: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

5

Animal Models of Cochlear Gene Therapy

•Hair Cell Regeneration

•Ototoxicity

•Spiral Ganglion Preservation

•Autoimmune Hearing Loss

•Genetic Hearing Loss

Animal Models of Cochlear Gene Therapy

•Hair Cell Regeneration

•Prevention of Ototoxicity

•Spiral Ganglion Preservation

•Autoimmune Hearing Loss

•Genetic Hearing Loss

Animal Models of Cochlear Gene Therapy

•Hair Cell Regeneration

•Prevention of Ototoxicity

•Spiral Ganglion Preservation

•Autoimmune Hearing Loss

•Genetic Hearing Loss

Math1 KO MiceBermingham et al, Science, 1999

• Failure of cochlear & vestibular HC’s to differentiate

• Cochlea/vestib endorgansotherwise nl

Page 6: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

6

Math1 gene transfer generates new cochlear

hair cells in mature guinea pigs in vivo

Kawamoto K et al,J Neurosci. 2003

Normal (Ad-perilymph)cochleaNormal (Ad-perilymph)cochlea

Kawamoto K et al

J Neurosci June 1 200323(1):4395-4400

Math1 Gene Transfer

Izumikawa et al. Nature Medicine 11, 271 - 276 (2005)Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals

Izumikawa et al. (2005)Auditory hair cell replacement and

hearing improvement by Atoh1 gene therapy

in deaf mammals

Page 7: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

7

Vestibular Hair Cell Regeneration and Restoration of Balance Function Induced by Math1 Gene Transfer.Staecker H et al. Otol Neurotol 2007

Control

Neomycin

Neomycin+Admath1

Neomycin+Admath1

Swim time

HVOR Recovery

Staecker H et al. Otol Neurotol 2007.

•Human Elongation factor 1-α promoter

•Math-1 insert

•In utero gene injection w/ electroporation

•Gain of function

Gubbels et al, 2008

“Acquired”

Genetic

Page 8: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

8

•GJB2=Connexin26 = autosomal recessive non-syndromic HL

•Lipid-mediated transfection of dominant allele causes HL

•Co-transfection of siRNA reduces hearing loss

• Single IP injection of ASO vs. defective pre-mRNA Ush 1C transcript

• Partially corrects splicing, increases protein expression, improves stereocilia organization, and rescues cochlear hair cells, vestibular function and low-frequency hearing

2013 Mar;19(3):345-50

2008

DAPI Synaptophysin VGLT3

VGLUT3 in the Organ of Corti

IHC OHC

TC

Seal et al, Neuron, 2008

Page 9: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

9

Am J Hum Genet 2008

•DFNA25 = autosomal-dominant progressive, high-frequency non-syndromic hearing loss

•2 unrelated families = a heterozygous missensemutation, c.632C/T was found to segregate with DFNA25 Neuron 75, 283–293, July 26, 2012

Page 10: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

11/8/2013

10

Delivery P10-12

Vglut3 ABR P40

01020304050

60708090

100

Click 8 KHz 16 KHz 32 KHz

Stimulus

AB

R T

hre

sho

ld d

B S

PL

Rescued KOWild TypeKO

Delivery P1-3

VGLUT3 Rescued MiceAcknowledgements

UCSF: Lustig LabOmar Akil, PhDSean AlemiKevin BurkeEdwards LabRobert Edwards

Ohio State UnivMatthew During,PhD

Univ PittsburghRebecca Seal, PhD

Johns Hopkins UnivElisabeth Glowatzki, PhD

Page 11: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

nature medicine volume 19 | number 3 | march 2013 345

Rescue of hearing and vestibular function by antisense oligonucleotides in a mouse model of human deafnessJennifer J Lentz1,6, Francine M Jodelka2,6, Anthony J Hinrich2,6, Kate E McCaffrey2, Hamilton E Farris1, Matthew J Spalitta1, Nicolas G Bazan3, Dominik M Duelli4, Frank Rigo5 & Michelle L Hastings2

Hearing impairment is the most common sensory disorder, with congenital hearing impairment present in approximately 1 in 1,000 newborns1. Hereditary deafness is often mediated by the improper development or degeneration of cochlear hair cells2. Until now, it was not known whether such congenital failures could be mitigated by therapeutic intervention3–5. Here we show that hearing and vestibular function can be rescued in a mouse model of human hereditary deafness. An antisense oligonucleotide (ASO) was used to correct defective pre-mRNA splicing of transcripts from the USH1C gene with the c.216G>A mutation, which causes human Usher syndrome, the leading genetic cause of combined deafness and blindness6,7. Treatment of neonatal mice with a single systemic dose of ASO partially corrects Ush1c c.216G>A splicing, increases protein expression, improves stereocilia organization in the cochlea, and rescues cochlear hair cells, vestibular function and low-frequency hearing in mice. These effects were sustained for several months, providing evidence that congenital deafness can be effectively overcome by treatment early in development to correct gene expression and demonstrating the therapeutic potential of ASOs in the treatment of deafness.

Usher syndrome is characterized by hearing impairment combined with retinitis pigmentosa and, in some cases, vestibular dysfunction. The fre-quency in the general population may be as high as 1 in 6,000 (ref. 8). Type 1 Usher syndrome is characterized by profound hearing impair-ment and vestibular dysfunction at birth and the development of retinitis pigmentosa in early adolescence. Approximately 6–8% of type 1 Usher syndrome cases are caused by mutations in the USH1C gene9, which encodes the protein harmonin. The USH1C216G>A (216A) mutation accounts for all cases of type 1 Usher syndrome in Acadian popula-tions9–11 and creates a cryptic 5ʹ splice site that is used preferentially over the authentic 5ʹ splice site of exon 3 (Fig. 1a), resulting in a frameshift and truncated harmonin protein12.

We used a mouse model of Usher syndrome based on the human 216A mutation13 to investigate a treatment for deafness and vestibular dysfunc-tion using ASOs (Supplementary Fig. 1) designed to redirect cryptic

splicing of 216A pre-mRNA to the authentic site (Fig. 1a). To screen for ASOs that block 216A cryptic splicing, we transfected a minigene expression plasmid composed of exons 2–4 and the intervening introns of human USH1C (wild type) or USH1C 216A into HeLa cells with 47 different individual ASOs surrounding the mutation and quantified splicing correction (Fig. 1b and Supplementary Table 1). Several ASOs blocked cryptic splicing and promoted correct splicing (Fig. 1b,c) in a dose-dependent manner (Fig. 1d). The ASOs also blocked cryptic splicing, promoted correct splicing of the endogenous Ush1c 216A gene transcript and increased harmonin protein expression in a mouse kidney cell line derived from mice homozygous for the Ush1c 216A mutation (216AA mice) (Supplementary Fig. 2). ASOs induced correct splicing in vivo after a series of intraperitoneal injections of 50 mg per kg body weight of ASO in adult 216AA mice (Fig. 1e). ASO-29 promoted the highest amount of correct splicing of the ASOs tested (Fig. 1e) and also corrected splicing and increased harmonin protein expression (Fig. 1f,g and Supplementary Fig. 3) in a dose-dependent manner.

216AA mice are deaf and have severe vestibular dysfunction, as indi-cated by their auditory brainstem response (ABR) and head-tossing and circling behavior13,14. Neonatal 216AA mice were injected intra-peritoneally with 300 mg per kg body weight of ASO-C or ASO-29 to test whether ASO-29 can correct vestibular and hearing defects. 216AA mice untreated or treated with a mismatched ASO (ASO-C) showed circling behavior, whereas 216AA mice treated with ASO-29 did not circle, and showed similar behavior to heterozygous (216GA) or wild-type (216GG) mice (Fig. 2a,b and Supplementary Video 1). We did not observe any circling behavior in mice treated with ASO-29 at postembryonic day 3 (P3), P5, P10 or P13, whereas 216AA mice treated on P16 showed circling behavior similar to untreated or ASO-C–treated 216AA mice (Fig. 2b). ASO-29–treated 216AA mice had no vestibular dysfunction at 6 months of age (Supplementary Fig. 4a). We performed trunk-curl, contact-righting and swim tests on 2- to 3-month-old and 6- to 9-month-old mice to further quantify ves-tibular function15. The younger and older 216AA mice all performed poorly in these tests, whereas 216AA mice treated with ASO-29 at P5 performed similarly to untreated or ASO-C–treated 216GA mice and showed no vestibular dysfunction (Supplementary Fig. 4b). Our

1Neuroscience Center and Department of Otorhinolaryngology & Biocommunications, Louisiana State University Health Sciences Center (LSUHSC), New Orleans, Louisiana, USA. 2Department of Cell Biology and Anatomy, Chicago Medical School, Rosalind Franklin University of Medicine and Science, North Chicago, Illinois, USA. 3Neuroscience Center and Department of Ophthalmology, LSUHSC, New Orleans, Louisiana, USA. 4Department of Cellular and Molecular Pharmacology, Chicago Medical School, Rosalind Franklin University of Medicine and Science, North Chicago, Illinois, USA. 5Isis Pharmaceuticals, Carlsbad, California, USA. 6These authors contributed equally to this work. Correspondence should be addressed to M.L.H. ([email protected]) or J.J.L. ([email protected]). Received 12 September 2012; accepted 28 January 2013; published online 4 February 2013; doi:10.1038/nm.3106

l e t t e r snp

201

3 N

atur

e A

mer

ica,

Inc.

All

right

s re

serv

ed.

Page 12: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

346 volume 19 | number 3 | march 2013 nature medicine

We recorded responses to different sound frequencies (8–32 kHz and broad band noise, BBN) at different intensities (18–90 dB SPL). Hearing thresholds represent the lowest sound intensity at which a recognizable ABR wave (neural response) is observed. We compared ABR thresholds in P30 216AA mice treated with ASO-29, treated and untreated 216GG wild-type and 216GA heterozygous mice, and 216AA mutant mice treated with ASO-C. 216GG and 216GA mice had thresholds typical of mice with normal hearing (Fig. 2c,d). Similar to untreated 216AA mice14, 216AA mice treated with ASO-C had abnormal or absent ABRs (Fig. 2c,d and Supplementary Fig. 6). 216AA mice treated between P3 and P5 with a single dose of ASO-29

results suggest that ASOs can prevent vestibular dysfunction associated with Usher syndrome in mice when delivered neonatally.

Treatment of 216AA mice with ASO-29 also rescued hearing. Startle responses to high-amplitude sound are similar in ASO-29–treated 216AA and either ASO-C–treated or untreated 216GA mice (Supplementary Video 2 and Supplementary Fig. 5). 216AA mice treated with ASO-C, however, showed neither an ini-tial startle response, defined as an ear twitch and rapid head and body movement, nor a subsequent freezing response after acoustic stimulus (Supplementary Video 2). We recorded auditory-evoked brainstem responses to quantitatively assess hearing function.

3 4

Correct splicing

216G>A

2

Cryptic splicing(frameshift)

135 aaTruncated protein

PDZ1

PDZ1

PDZ1

a

b

c

PDZ2

PDZ2

PDZ2

CC1

CC1

CC1

CC2

PDZ3

PST PDZ3

Full-length harmonin

552 aa

899 aa

403 aa

USH1C

Exon 2 Exon 3 Exon 4216A

12 3 4 5–8 9–12 13–16 17–20 21–24 25–28 29–32 33–36 37–40 41–44

454647

UnsplicedCorrectCryptic

Skip

ASO 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 C C

10 1 1 0 0 0 0 0 0 0 0 0 0 1 0 0 00 0 1 0 1 1 11 1 4 2 2 0 0 0 0 0 0 0 0 1 2 4 2 1 0 1 0 0 0 0

13 6 10 32 35 64 24 35 42 12 37 26 22 30 64 16 58 5 4 10 22 17 22 24 15 2 12 3 3 20 99 99 100 91 98 100

100 98 96 95 97 98 98 98 94 98 98 98 98

Correct (%)Cryptic (%)

ASO-49ASO-48ASO-29ASO-28USH1C

Authentic 5′ splice site216ACryptic 5′ splice site

MinigeneASO(nM)

28 29 48 49

USH1C 216A

UnsplicedCorrectCryptic

Skip

ASO(nM)

28 29 48 49

5 10 20 40 80 5 10 20 40 80 5 10 20 40 80 5 10 20 80

5 10 20 40 80 5 10 20 40 80 5 10 20 40 80 5 10 20 80

Correct Cryptic

020406080

100

Splic

ing

(%)

Mice 216AAASO

28 29 48 49―

CorrectCryptic

Skip

28 29 48 49―ASO0

10

20

30

Corre

ct s

plici

ng (%

)

―ASO ASO-29

CorrectCryptic

Skip

(mg per kg BW) 0 50 100 200

0 50 100 2000

102030

Corre

ct

splic

ing

(%)

(mg per kg BW)

―ASO ASO-29(mg per kg BW) 0 50 100

b

a

c

Harmonin

β-actin17

2433405572

100135

MW (kDa)

******

**

**

*

a

b

c

d

e

f g

Figure 1 Correction of USH1C 216A splicing using ASOs. (a) Gene structure of USH1C exons 2–4 and RNA splicing and protein products. Boxes represent exons and lines are introns. Diagonal lines indicate splicing. The locations of the 216A mutation and the cryptic splice site are labeled. aa, amino acids; PDZ, PSD95-Dlg1-zo1-domain; CC, coiled-coil domain; PST, proline-serine-threonine–rich domain. (b) Top, diagram of ASOs tested below, mapped to their position of complementarity on USH1C. Bottom, radioactive RT-PCR of RNA isolated from HeLa cells transfected with a USH1C 216A minigene and the indicated ASO at a final concentration of 50 nM. RNA spliced forms are labeled. ‘Unspliced’ refers to transcripts with intron 3 retained, and ‘skip’ indicates exon 3 skipping. (Below) Quantification of the percentage splicing in graph is calculated as: Correct (%) = [(correct/(correct + cryptic + skip)] × 100, and Cryptic (%) = [(cryptic/(correct + cryptic + skip)] × 100. ‘C’ indicates mock-treated control. (c) Sequence and USH1C target region of ASOs. Exonic and intronic sequences are in capital and lower-case letters, respectively. (d) Analysis of USH1C 216A minigene transcript splicing in HeLa cells treated with increasing concentrations of the indicated ASOs from c. Quantification of percentage correct and percentage cryptic splicing is shown in graph (below). (e) RT-PCR analysis of RNA isolated from kidneys of adult 216AA mice 24 h after final injection of 50 mg per kg body weight (BW) of different ASOs. Samples from three individual mice are shown. Ush1c spliced products are indicated and quantified in graph (below) as described above. Error bars represent s.e.m. (*P ≤ 0.05, **P ≤ 0.01, n = 3, two-tailed Student’s t-test compared to vehicle treatment). (f) RT-PCR analysis of RNA isolated from kidneys of adult Ush1c 216AA mice treated with different doses of ASO-29. Samples from three individual mice are shown. Ush1c spliced products are indicated and quantified as described above. Error bars represent s.e.m. (***P ≤ 0.001, n = 3, two-tailed Student’s t-test compared to vehicle). (g) Immunoblot analysis of harmonin protein in lysates from the kidneys of adult 216AA mice analyzed in f. Blots were also probed with a b-actin–specific antibody for a loading reference. MW, molecular weight.

l e t t e r snp

201

3 N

atur

e A

mer

ica,

Inc.

All

right

s re

serv

ed.

Page 13: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

nature medicine volume 19 | number 3 | march 2013 347

216AA mice (P ≤ 0.05) (Supplementary Fig. 9a,b). These results sug-gest that mice injected with a single ASO treatment early in life can hear at 6 months of age, indicating a long-term, if slowly declining, therapeutic rescue of hearing.

To determine the effect of ASOs on Ush1c mRNA splicing and harmo-nin expression, we analyzed the cochleae from mice injected at P5 with ASO-29 or ASO-C. At P30, we observed a low amount of correct exon 3 splicing in the ASO-29–treated 216AA mice (Fig. 3a). Correct splicing peaked at P120, and expression of the full-length mRNA transcript at P180 was similar to that at P30, indicating that the effect of the ASO on splicing is stable and corresponds with ABR results at this time point (Supplementary Fig. 9c). Harmonin protein abundance was also higher in cochleae from ASO-29–treated 216AA mice compared to 216AA mice treated with ASO-C and similar to harmonin levels at P32–P35 in the cochleae of control 216GA mice (Fig. 3b and Supplementary Fig. 10).

The harmonin b isoform (Fig. 1a) localizes to the developing and mature stereocilia bundle of cochlear hair cells16–19, where it is hypothe-sized to scaffold the molecular components of the mechanotransduction machinery20. We examined the expression and localization of harmonin b in microdissected organs of Corti from P30 216AA mice injected at P5 with either ASO-29 or ASO-C and untreated 216GA heterozygous lit-termates. Harmonin b was abundantly expressed in the tips of outer hair cell stereocilia bundles of 216GA mice, whereas the 216AA mice had less expression and it was mislocalized in the atypical bundles (Fig. 3c,d).

had recognizable waveforms and near-normal thresholds to BBN and 8- and 16-kHz pure tones when compared to 216GG and 216GA control mice (Fig. 2c,d and Supplementary Fig. 6). The thresholds of ASO-29–treated 216AA mice at 32 kHz were not significantly dif-ferent than those of control ASO-C–treated 216AA mice, indicating that treatment was not effective at rescuing high-frequency hearing (Fig. 2d). These data show rescue of low- and mid-frequency hear-ing at P30 in ASO-29–treated 216AA mice. 216AA mice treated with ASO-29 at P10 had significantly higher thresholds than those treated at P3–P5 but significantly lower thresholds in response to BBN and an 8-kHz tone than untreated mutants or mutants treated with ASO-C (P ≤ 0.05) (Fig. 2c,d and Supplementary Fig. 6), implying a develop-mental window of therapeutic efficacy in mice. At 2 months of age, 216AA mice treated at P3–P5 had ABR thresholds to BBN and 8 kHz and 16 kHz, but not 32 kHz, equivalent to 216GG and 216GA mice (P ≤ 0.05, Fig. 2e and Supplementary Fig. 7a–d). At 3 months, there was no significant difference in ABR thresholds at 8 kHz between control 216GA heterozygous mice and 216AA mutant mice treated with ASO-29 at P3–P5 (P ≤ 0.05), but the data show some loss of sen-sitivity in ASO-29–treated 216AA mice to 16 kHz and BBN (Fig. 2f and Supplementary Fig. 8a–d). At 6 months of age, there were signifi-cant differences in ABR thresholds between control 216GA or ASO-C–treated 216GG mice and ASO-29–treated 216AA mice at all frequencies (P ≤ 0.05), though ASO-29–treated 216AA mice showed ABR thresh-olds that were significantly lower than those of ASO-C–treated

a b

c

d e f

Mutant (216AA)ASO-C

Mutant (216AA)ASO-29

Heterozygote (216GA)ASO-C

*** *** *** *** NS ***

***NS NS NS NS

AA AA AA AA AA AA AA GA GA GG– P3-P16 P3 P5 P10 P13 P16 P5 – –

ASO-29 – –ASO-C–

Rot

atio

ns p

er 1

20 s

0

10

20

30

40

50

Treatment

Treatment day

Ush1c216 genotype

Mutant (216AA)ASO-C, P5

Mutant (216AA)ASO-29, P5

Mutant (216AA)ASO-29, P10

Heterozygote (216GA)

8 kH

z2 uV

1 m

90 dB SPL78 dB SPL66 dB SPL54 dB SPL48 dB SPL42 dB SPL30 dB SPL24 dB SPL

1 month 2 months 3 months

AA, ASO-C, P5 AA, ASO-29, P5 AA, ASO-29, P10 GG/GA, controlSound frequency (kHz)

Thre

shol

d (d

B SP

L)

5 8 16 32 BBN0

20

406080

100

5 8 16 32 BBN0

20

406080

100

5 8 16 32 BBN0

20

406080

100

n = 39 n = 21 n = 9 n = 32 n = 7 n = 4 n = 8 n = 17 n = 30 n = 5

Figure 2 ASOs correct vestibular function and rescue hearing in 216AA mice. (a) Representative open-field pathway traces (120 s) from a P22 mouse in each group are shown. (b) Quantification of the number of rotations in 120 s from P22–P35 mice. Error bars represent s.e.m., the number (n) of mice analyzed is indicated within the individual bars. Significance (***P ≤ 0.001 or not significant, NS) was calculated using one-way analysis of variance (ANOVA) and Tukey-Kramer post test. (c) Representative ABR waveforms at 8-kHz stimulus from a 216AA mouse injected at P5 with ASO-C (left), a 216AA mouse injected at P5 or P10 with ASO-29 (middle) and a 216GA mouse (right). Colored lines (red = AA, ASO-C; blue = AA, ASO-29 injected at P5; orange = AA, ASO-29 injected at P10; gray = GA control) represent thresholds detected. (d–f) Average ABR thresholds (dB SPL) to pure tones ranging from 8 to 32 kHz or BBN in 216AA mutant (AA) and 216 GG wild-type (GG) or 216 GA heterozygous (GA) mice at 1 month of age (n = 11, 8, 5 and 11 for 216AA, ASO-C; 216AA, ASO-29 at P5; 216AA, ASO-29 at P10 and 216GG/GA, respectively (wild-type and heterozygotes were grouped together because we found no difference between them)) (d), 2 months of age (n = 4, 6 and 5 for 216AA, ASO-C; 216AA, ASO-29; and 216GG/GA, respectively) (e) and 3 months of age (n = 3, 4, and 4 for 216AA, ASO-C; 216AA, ASO-29; and 216GG or GA, respectively) (f). Error bars represent s.e.m. * indicates a significant difference between ASO-29–treated 216AA and ASO-C–treated 216AA mice, and # indicates a significant difference between ASO-29–treated 216AA and 216GA control mice (P ≤ 0.05; two-way ANOVA with Tukey-Kramer post test).

l e t t e r snp

201

3 N

atur

e A

mer

ica,

Inc.

All

right

s re

serv

ed.

Page 14: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

348 volume 19 | number 3 | march 2013 nature medicine

stereocilia bundles with typical ‘U’ or ‘W’ bundle shapes (Fig. 4a,b). The 216AA mice treated with ASO-29 had significantly fewer atypical bun-dles in the regions that detect 8 (0.8–1.2 mm) and 16 kHz (1.8–2.2 mm) (P ≤ 0.05) but not in the region detecting 32 kHz (3.8–4.2 mm) (Fig. 4c). This pattern of anatomical stereocilia rescue is consistent with the ABR results showing a rescue of hearing in the 8- and 16-kHz range and less robust rescue in the 32-kHz range. Together, our results indicate a change in the bundle structure and number of hair cells at the apical–mid regions in ASO-29–treated 216AA mice, providing an anatomical correlate to function.

Our study demonstrates that a human disease–causing mutation can be corrected to treat deafness and vestibular dysfunction in a mouse model. Treatment during the critical hair cell developmental period is probably necessary and perhaps sufficient for long-term rescue of hearing. Notably, a modest correction of Ush1c mRNA splic-ing and expression of full-length harmonin is sufficient to rescue hearing in mice for 6 months. This enduring effect of the ASOs is consistent with the duration of action of ASOs observed in other mouse models of disease such as spinal muscular atrophy23 and myo-tonic dystrophy24.

ASO-29–treated 216AA mice had elevated harmonin b expression with localization at the tips of stereocilia similar to that in the 216GA mice (Fig. 3c,d). These results suggest proper localization of harmonin with ASO treatment.

Frequency place-mapping in the mouse cochlea21,22 suggests that the region corresponding to 8–16 kHz (the frequencies most robustly rescued by ASO-29) is located approximately 1–2 mm from the apex tip (Fig. 3f). To assess the relationship between hair cell number and ABR threshold, we counted hair cells labeled with parvalbumin. At P35, 216AA mice had significant outer hair cell loss from approxi-mately 0.8–2.0 mm from the apex, corresponding to hearing at 6–20 kHz (P ≤ 0.05) (Fig. 3g,h). In contrast, the number of outer hair cells in this region of 216AA mice treated with ASO-29 at P3–P5 did not differ from that in 216GA mice, which is consistent with rescued physi-ological function (Fig. 3g,h).

We also assessed changes in hair cell morphology that may reflect the rescue of hearing in the regions of the cochlea sensitive to 8 and 16 kHz. By P35, 216AA mice had significant hair cell loss in this region (Fig. 3); therefore, we analyzed subcellular structures in P18 216AA and 216GA mice before the loss of hair cells. We quantified the number of

Heterozygote (216GA)ASO-C

Mutant (216AA)ASO-29

Mutant (216AA)ASO-C

GA, ASO-C AA, ASO-29 AA, ASO-C

IHCs OHCs

Distance from apex tip (mm)

Avg.

no.

hai

r cel

lspe

r 100

μm

20

25

30

35

40

0 0.5 1.0 1.5 2.0 0 0.5 1.0 1.5 2.00

25

50

75

100

125

Heterozygote (216GA)ASO-C

Mutant (216AA)ASO-29

Mutant (216AA)ASO-C

Dis

tanc

e fro

m a

pex

tip (m

m)

1.8–

2.0

0.8–

1.0

16 kHz

1.0 mm

8 kHz

OHCs

IHCsApex tip

ParvalbuminDAPI

b

a

c

Harm

oninβ-actin

172433405572

100135

Mouse MW (kDa)TreatmentGenotype

1 2 3 4 5 6 7 8 9ASO-29— —

AA GA

CorrectCryptic

Skip

Correct (%)

Mouse

TreatmentGenotype

1 2 3 4 5 6 7 8 9

ASO-29— —AA GA GG

0.25± 0.08n = 10

1.6± 0.68n = 6

41± 13.1n = 4

100± 0

n = 3

a c

bd

e

f g

h

Figure 3 ASO-29 treatment corrects mRNA splicing and harmonin protein expression and prevents cochlear hair cell loss in 216AA mice. (a) RT-PCR analysis of cochlear RNA isolated at P32–P35 from mice treated with control (ASO-C) or ASO-29 at P3–P5. Spliced products are labeled. (b) Immunoblot analysis of harmonin expression in cochlea isolated at P32–35 from mice that were treated at P5. Different isoforms of harmonin expressed from Ush1c are labeled. Blots were also probed with a b-actin–specific antibody for a loading reference. (c) Immunofluorescence staining of harmonin b (green) and F-actin (red, phalloidin) in outer hair cell (OHC) bundles in the region of the basilar membrane that corresponds to hearing at 8 kHz (0.8–1.5 mm from apex tip). Images are from 216GA mice (left), 216AA mice treated with ASO-29 at P5 (middle) or 216AA mice treated with ASO-C at P5 (right). Scale bar, 3 mm. (d) Digital magnification (2× zoom) of images shown in c. Scale bar, 2 mm. (e) Immunofluorescence image of a primary antibody isotype control from a 216GA mouse taken from a similar OHC bundle location. (f) Immunofluorescence image of the regions of the basilar membrane that are represented in g and h. Hair cells are labeled with parvalbumin (red), and nuclei are counterstained with DAPI (blue). Scale bar, 50 mm. IHCs, inner hair cells. (g) Cochleogram showing inner (left) and outer (right) hair cell counts from regions progressively distant from the apex tip. Error bars represent the s.e.m. (*P ≤ 0.05; n = 3 mice). At least 100 cells from each experimental group were evaluated for each region. (h) Immunofluorescence images of representative regions along the basilar membrane 0.8–1.0 mm (top) or 1.8–2.0 mm (bottom) from the extreme apex from P35 216GA mice treated with ASO-C at P5 (left), 216AA mice treated with ASO-29 at P5 (middle) or 216AA mice treated with ASO-C at P5 (right). Scale bar, 20 mm.

l e t t e r snp

201

3 N

atur

e A

mer

ica,

Inc.

All

right

s re

serv

ed.

Page 15: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

nature medicine volume 19 | number 3 | march 2013 349

occurs at the base of the cochlea, this result may suggest that Ush1c is expressed tonotopically during development and that splicing correction at P5 benefits only the 30. apical-mid regions of the cochlea. The develop-ment of the ear and hearing in humans occurs in utero27. Thus, treatment in humans would probably require delivery to the fetus via approaches such as intrauterine transfusion28.

Individuals with Usher syndrome present with both hearing and vision loss, and the correction of one of these sensory deficits may have a considerable positive impact. Although the retinitis pigmentosa associated with Usher syndrome is recapitulated in the 216AA mice, their retinal cell loss occurs later, at approximately 1 year of life14. Thus, our analysis of vision in these mice will require further investigation at later time points. The rescue of hearing in this study offers a model for studying the development of hearing and vestibular function and for developing approaches to correct these processes when they are impaired.

METHODSMethods and any associated references are available in the online ver-sion of the paper.

Note: Supplementary information is available in the online version of the paper.

ACKNOWLEDGMENTSWe gratefully acknowledge support from the Hearing Health Foundation, Midwest Eye-Banks, the National Organization for Hearing Research Foundation, Capita Foundation and the US National Institutes of Health. We thank D. Cunningham and E. Rubel for assistance with scanning electron microscopy analysis; A. Rosenkranz, R. Marr and M. Oblinger for use of equipment; J. Huang for assistance with open-field analysis, L. Ochoa for assistance with ABR analysis computer support, G. MacDonald for assistance with confocal imaging and deconvolution analysis, U. Wolfrum (Johannes Gutenberg University of Mainz) for harmonin b–specific antibodies, H. Thompson for statistical analysis, and A. Case, B. Keats and M. Havens for discussions and comments on the manuscript.

AUTHOR CONTRIBUTIONSThe project was conceived of by M.L.H. Experiments were designed and performed by A.J.H., F.M.J., J.J.L., K.E.M., M.L.H. and D.M.D., and were analyzed by M.L.H., J.J.L., F.R., F.M.J., A.J.H. and K.E.M. Animal work was carried out by M.L.H., D.M.D., K.E.M., A.J.H., F.M.J., J.J.L. and M.J.S. Molecular experiments were performed by A.J.H., F.M.J., K.E.M. and M.L.H. J.J.L. carried out the immunofluorescence analysis. J.J.L., M.J.S. and H.E.F. performed auditory brainstem response experiments, and J.J.L. and N.G.B. interpreted the results. M.L.H. and J.J.L. wrote the paper.

COMPETING FINANCIAL INTERESTSThe authors declare competing financial interests: details are available in the online version of the paper.

Reprints and permissions information is available online at http://www.nature.com/reprints/index.html.

1. Morton, C.C. & Nance, W.E. Newborn hearing screening—a silent revolution. N. Engl. J. Med. 354, 2151–2164 (2006).

2. Dror, A.A. & Avraham, K.B. Hearing loss: mechanisms revealed by genetics and cell biology. Annu. Rev. Genet. 43, 411–437 (2009).

3. Conde de Felipe, M.M., Feijoo Redondo, A.F., Garcia-Sancho, J., Schimmang, T. & Alonso, M.B. Cell- and gene-therapy approaches to inner ear repair. Histol. Histopathol. 26, 923–940 (2011).

4. Di Domenico, M. et al. Towards gene therapy for deafness. J. Cell. Physiol. 226, 2494–2499 (2011).

5. Bermingham-McDonogh, O. & Reh, T.A. Regulated reprogramming in the regeneration of sensory receptor cells. Neuron 71, 389–405 (2011).

6. Bitner-Glindzicz, M. et al. A recessive contiguous gene deletion causing infantile hyper-insulinism, enteropathy and deafness identifies the Usher type 1C gene. Nat. Genet. 26, 56–60 (2000).

7. Verpy, E. et al. A defect in Harmonin, a PDZ domain-containing protein expressed in the inner ear sensory hair cells, underlies Usher syndrome type 1C. Nat. Genet. 26, 51–55 (2000).

8. Kimberling, W.J. et al. Frequency of Usher syndrome in two pediatric populations: Implications for genetic screening of deaf and hard of hearing children. Genet. Med. 12, 512–516 (2010).

The rescue of hearing in mice by ASO-29 treatment demonstrates that deafness can be treated if intervention occurs early in development. Treatment at P10 corrects vestibular function and partially restores hear-ing, whereas treatment at P3–P5 rescues vestibular function and hearing with ABRs comparable to those of wild-type mice (Supplementary Fig. 1). Although Ush1c is expressed as early as embryonic day 15 in mice25,26, there is a peak of developmental expression in the cochlea that occurs after P4 and before P16 (https://shield.hms.harvard.edu/viewgene.html?gene=Ush1c). Our results are consistent with this expression pat-tern, suggesting that high expression before P5 is not required for the development of low- and mid-frequency hearing, but expression between P5 and P10 may be crucial. Hearing at high frequencies (32 kHz) is not rescued to the same level as that at the lower frequencies, and the rescue is more transient (Fig. 2). Because detection of high-frequency sound

Heterozygote(216GA)

Mutant(216AA)ASO-29

Mutant(216AA)ASO-C

0.8–

1.2

mm

1.8–

2.2

mm

3.8–

4.2

mm

Dis

tanc

e fro

m a

pex

tip

16 kHz

Apextip

8 kHz32 kHz1 mm

3 mm

4 mm

2 mm

100 μm

AA, ASO-29 AA, ASO-C

10 100Frequency (kHz)At

ypic

al s

tere

ocilia

bun

dles

(%)

0

20

40

60

80

100

b

a

c

Figure 4 Restoration of hair cell stereocilia bundle shape in mice. (a) Scanning electron micrographs of outer hair cell bundles from P18 216GA and 216AA mice treated at P5 with ASO-C or ASO-29. Distance (mm) was measured from apex tip. Scale bars represent 1 mm and 10 mm for the high- and low-magnification images, respectively. (b) Scanning electron micrograph illustrating the regions of the cochlea that are represented in a. Scale bar, 100 mm. (c) Quantification of atypical bundles shown as a percentage of total cells counted at different positions along the basilar membrane in 216AA mice treated either with ASO-29 (blue line) or ASO-C (red line) (*P ≤ 0.05, **P ≤ 0.005; two-tailed unpaired t-test; n = 3 or 4 mice per region). Error bars represent s.e.m. At least 200 cells from each experimental group were evaluated with at least 60 hair cells from each region. 216GA control mice have no atypical bundles (data not shown).

l e t t e r snp

201

3 N

atur

e A

mer

ica,

Inc.

All

right

s re

serv

ed.

Page 16: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

350 volume 19 | number 3 | march 2013 nature medicine

links of the hair bundle in its cohesion, orientation and differential growth. Development 135, 1427–1437 (2008).

19. Boëda, B. et al. Myosin VIIa, harmonin and cadherin 23, three Usher I gene products that cooperate to shape the sensory hair cell bundle. EMBO J. 21, 6689–6699 (2002).

20. Peng, A.W., Salles, F.T., Pan, B. & Ricci, A.J. Integrating the biophysical and molecular mechanisms of auditory hair cell mechanotransduction. Nat. Commun. 2, 523 (2011).

21. Müller, M., von Hunerbein, K., Hoidis, S. & Smolders, J.W. A physiological place-frequency map of the cochlea in the CBA/J mouse. Hear. Res. 202, 63–73 (2005).

22. Greenwood, D.D. A cochlear frequency-position function for several species—29 years later. J. Acoust. Soc. Am. 87, 2592–2605 (1990).

23. Hua, Y. et al. Peripheral SMN restoration is essential for long-term rescue of a severe spinal muscular atrophy mouse model. Nature 478, 123–126 (2011).

24. Wheeler, T.M. et al. Targeting nuclear RNA for in vivo correction of myotonic dystrophy. Nature 488, 111–115 (2012).

25. El-Amraoui, A. & Petit, C. Usher I syndrome: unravelling the mechanisms that underlie the cohesion of the growing hair bundle in inner ear sensory cells. J. Cell Sci. 118, 4593–4603 (2005).

26. Petit, C. & Richardson, G.P. Linking genes underlying deafness to hair-bundle develop-ment and function. Nat. Neurosci. 12, 703–710 (2009).

27. Hall, J.W. III. Development of the ear and hearing. J. Perinatol. 20, S12–S20 (2000).28. Uhlmann, R.A., Taylor, M., Meyer, N.L. & Mari, G. Fetal transfusion: the spectrum of

clinical research in the past year. Curr. Opin. Obstet. Gynecol. 22, 155–158 (2010).

9. Ouyang, X.M. et al. Characterization of Usher syndrome type I gene mutations in an Usher syndrome patient population. Hum. Genet. 116, 292–299 (2005).

10. Ebermann, I. et al. Deafblindness in French Canadians from Quebec: a predominant founder mutation in the USH1C gene provides the first genetic link with the Acadian population. Genome Biol. 8, R47 (2007).

11. Ouyang, X.M. et al. USH1C: a rare cause of USH1 in a non-Acadian population and a founder effect of the Acadian allele. Clin. Genet. 63, 150–153 (2003).

12. Lentz, J. et al. The USH1C 216G→A splice-site mutation results in a 35-base-pair deletion. Hum. Genet. 116, 225–227 (2005).

13. Lentz, J., Pan, F., Ng, S.S., Deininger, P. & Keats, B. Ush1c216A knock-in mouse survives Katrina. Mutat. Res. 616, 139–144 (2007).

14. Lentz, J.J. et al. Deafness and retinal degeneration in a novel USH1C knock-in mouse model. Dev. Neurobiol. 70, 253–267 (2010).

15. Hardisty-Hughes, R.E., Parker, A. & Brown, S.D. A hearing and vestibular phenotyping pipeline to identify mouse mutants with hearing impairment. Nat. Protoc. 5, 177–190 (2010).

16. Michalski, N. et al. Harmonin-b, an actin-binding scaffold protein, is involved in the adaptation of mechanoelectrical transduction by sensory hair cells. Pflugers Arch. 459, 115–130 (2009).

17. Grillet, N. et al. Harmonin mutations cause mechanotransduction defects in cochlear hair cells. Neuron 62, 375–387 (2009).

18. Lefèvre, G. et al. A core cochlear phenotype in USH1 mouse mutants implicates fibrous

l e t t e r snp

201

3 N

atur

e A

mer

ica,

Inc.

All

right

s re

serv

ed.

Page 17: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

tests were performed by placing mice in a tub of room-temperature water and observing their swimming behavior for 10 s. Contact-righting reflex testing was performed by placing the mouse into a closed clear tube or box and measuring the time it took to right when turned upside down. The trunk-curl test was performed by holding the tail and observing whether the mouse reached for a nearby surface or curled toward the base of its tail.

Auditory-evoked brain stem response. Hearing thresholds of treated and untreated Ush1c 216GG, 216GA and 216AA mice were measured by auditory-evoked brain stem response (ABR). Mice were anesthetized (intraperitoneal ket-amine, 100 mg per kg body weight; xylazine, 6 mg per kg body weight), and body temperature was maintained near 38 °C with a heat pad. All recordings were con-ducted in a sound-proof room. Stimuli consisted of 5-ms pulses of broad-band noise, 8, 16 and 32 kHz, with 0.5-ms linear ramps. Although these tone stimuli encompass low, medium and high regions of mouse spectral sensitivity, BBN was included to confirm that responses are representative of the whole cochlear response. The stimuli were broadcast through a Motorola piezoelectric speaker (model no. 15D87141E02) fitted with a plastic funnel and 2-mm-diameter tubing over the speaker front, producing an acoustic wave guide which was positioned in the external meatus approximately 0.5 cm from the tympanum. Using continuous tones, stimulus amplitude was calibrated at the end of the tubing with a Bruel and Kjaer 2610 measuring amplifier (fast, linear weighting), 4135 microphone (grid on) and 4230 pistonphone calibrator. All stimulus amplitudes were dB SPL (re. 20 mPa). Total harmonic distortion was –40 dB (Hewlet Packard 3562A Signal Analyzer). Stimuli were generated (195 kHz srate) and responses digitized (10 kHz srate) using TDT System III hardware and software (BioSig). ABRs were recorded with a 30-gauge subdermal steel electrode placed subcutaneously behind the left ear, with indifferent and ground electrodes (steel wire, 30-gauge) placed subcutaneously at the vertex and hind limbs, respectively. After amplifi-cation (60 dB, Grass P5 AC), filtering (0.3 Hz–1 kHz; TDT PF1), and averaging (n = 600–1,024), thresholds (± 6 dB) were determined by eye as the minimum stimulus amplitude that produced an ABR wave pattern similar to that produced for the highest intensity stimulus (90 dB).

Scanning electron microscopy. The scanning electron microscopy analysis was performed as has been previously described14. Specifically, intralabyrinthine per-fusion with fixative (2.5% glutaraldehyde/1% paraformaldehyde/1.5% sucrose in 0.12 M phosphate buffer (pH 7.4)) was performed on whole cochleae dis-sected from mice at P18. Cochleae were post-fixed by immersion in the same fixative for 1 d at 4 °C with gentle rotation followed by three washes in 0.12 M phosphate-buffered saline (PBS) and stored for 1 week at 4 °C. Cochleae were next fixed in 1% OsO4 in PBS for 40 min and washed in PBS. Specimens were then serially dehydrated in ethanol, dried in a critical point drier (Autosamdri-814, Tousinis Research Corporation) and mounted on aluminum stubs. The bony capsule of the cochlea, spiral ligament, stria vascularis and Reissner’s membrane were removed, and the whole organ of Corti was exposed with fine dissecting instruments. Specimens were coated in gold/palladium with a Hummer VIA sputter coater (Anatech) and viewed on a JEOL JSM 6300 F scanning electron microscope. At least three individual animals representative of each experimental paradigm were analyzed.

The cochlear place-frequency map relating distance from cochlear apex and frequency is based on a tonotopic map of mice with an average basilar membrane distance from apex to base of 5.13 mm21. Distances were calculated using the equation: d = 156.5 – 82.5 × log(f); d is the normalized distance from the base (%) and f the frequency in kHz21.

Immunofluorescence. Fluorescent labeling of microdissected preparations of the organ of Corti was used to study the hair cells of 1-month-old treated and untreated mutant and control mice as described previously14,31. Briefly, cochleae were isolated from the auditory bulla, and a small opening was created in the apex. The stapes was removed from the oval window, and the cochleae were gen-tly perfused with 2% paraformaldehyde in 0.1 M phosphate buffer, pH 7.4 and post-fixed by immersion for 2 h at 4 °C with gentle rocking. Tissues were washed twice with PBS following fixation and processed for immunohistochemistry. For harmonin analysis, the tectorial membrane was removed with a fine forceps, and the stria vascularis was trimmed. Tissues were blocked for 1 h at room tempera-ture or overnight at 4 °C (harmonin analysis) in a blocking solution consisting of 10% normal donkey serum, 0.5% bovine serum albumin, 0.1% Triton X–100 and

ONLINE METHODSOligonucleotide synthesis. The synthesis and purification of all 2ʹO–methoxyethyl– modified oligonucleotides with phosphorothioate backbone and all 5–methyl cytosines, was performed as described29. The oligonucleotides were dissolved in 0.9% saline and stored at –20 °C. Sequences are shown in Supplementary Table 1.

Plasmids. The minigene expression plasmids pCI–Ush1C_216G and pCI–Ush1C_216A were constructed by amplifying genomic DNA from lymphoblast cell lines derived from an individual with Usher syndrome homozygous for the USH1C c.216G>A mutation USH1C.216AA (GM09458, Coriell Institute) or a healthy individual (GM09456, Coriell Institute). PCR primers specific for the 5ʹ end of exon 2 with restriction sites for XhoI and for the 3ʹ end of exon 4 with a restriction site for NotI at the 3ʹ end were used to amplify by PCR the USH1C 216A minigene fragment. The PCR product was purified and digested with XhoI and NotI and ligated into the expression plasmid pCI expression vector (Promega) digested with the same restriction enzymes.

Cell culture. Plasmids (1 mg) expressing a minigene of human USH1C 216A exons 2–4 and ASOs (50 nM final concentration) were transfected into HeLa cells using Lipofectamine 2000 (Life Technologies). Forty-eight hours after transfec-tion, RNA was isolated using Trizol reagent (Life Technologies) and analyzed by radioactive RT-PCR with the primers pCI FwdB and pCI Rev (Supplementary Table 1) to plasmid sequences flanking exon 2 and exon 4.

Mice. Ush1c 216A knock-in mice were obtained from LSUHSC13 and bred and treated at Rosalind Franklin University of Medicine and Science (RFUMS). All procedures met the NIH guidelines for the care and use of laboratory animals and were approved by the Institutional Animal Care and Use Committees at RFUMS and LSUHSC. Mice were genotyped using ear punch tissue and PCR as described previously14. For all studies, both male and female mice were used in approximately equal proportions. For studies in adult mice (Fig. 1e,f), homozy-gous 216AA mice (2–4 months of age) were injected intraperitoneally with the indicated ASO at indicated dose twice a week for 2 weeks (4 doses). RNA was isolated from different tissues 24 hrs after final injection using Trizol reagent (Life Technologies) and analyzed by radioactive RT-PCR using primers musU-SH1Cex2F and musUSH1Cex5F (Supplementary Table 1) of the Ush1c 216A transgene. Products were separated on a 6% nondenaturing polyacrylamide gel and quantified using a Typhoon 9400 phosphorimager (GE Healthcare). For studies in neonate mice, pups were injected with 300 mg per kg body weight of 2ʹMOE ASOs at different ages, post-natal day 3–16 (P3–P16), as indicated, by intraperitoneal injection. After ABR analysis, mice were killed and tissues were collected. For ABR analysis, mice were shipped to LSUHSC 2–3 weeks after treat-ment. ABR was carried out at least 3 d after mice arrived.

Splicing and protein analysis. Retinae and inner ears were isolated, and cochleae and vestibules were separated and immediately frozen in liquid nitrogen or stored in Trizol reagent. For immunoblot analysis, proteins were obtained from homogenization in a modified RIPA buffer30 or isolated from Trizol reagent (Life Technologies) according to the manufacturer’s instructions. Proteins were separated on 4–15% Tris-glycine gradient gels, transferred to membrane and probed with USH1C- (20900002, Novus Biologicals) or b-actin– (Sigma-Aldrich) specific antibodies. Blots were quantified using ImageJ v1.45s software (NIH). RNA was isolated from different tissues using Trizol reagent (Life Technologies) and analyzed by radioactive RT-PCR using primers musUSH1Cex2F and musU-SH1Cex5F of the Ush1c 216A transgene. Briefly, 0.25–1 mg of RNA was reverse transcribed using GoScript Reverse Transcriptase (Promega, Fitchburg, WI), and 1 ml of cDNA was used in PCR reactions with GoTaq Green (Promega) supple-mented with primers and 0.1–.25 ml of a-32P-dCTP. Products were separated on a 6% nondenaturing polyacrylamide gel and quantified using a Typhoon 9400 phosphorimager (GE Healthcare).

Behavioral analysis. Behavioral tests were performed according to previously established protocols15. Investigators were blinded except in cases where the phenotype made the treatment/genotype status obvious. To quantify circling behavior, mice were placed in an open-field chamber, and behavior was analyzed using ANY-maze behavioral tracking software (Stoelting Co). Ear-twitch, startle and freezing behavior in response to a high-amplitude sound was measured by observing mouse activity following a short whistle (Supplementary Fig. 4). Swim

nature medicine doi :10.1038/nm.3106

npg

© 2

013

Nat

ure

Am

eric

a, In

c. A

ll rig

hts

rese

rved

.

Page 18: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

0.03% saponin in PBS to reduce nonspecific binding of primary and secondary antibodies. Primary antibody incubations were then performed at 4 °C in PBS containing 5% normal donkey serum, 0.05% bovine serum albumin, 0.1% Triton X-100 and 0.03% saponin in PBS. For counting cells, a mouse monoclonal anti-parvalbumin antibody (parv19, P3088, Sigma, 1:250) was used to label cochlear hair cells32. To analyze harmonin b expression, polyclonal rabbit anti-harmonin antibodies specific to isoform b (gift from U. Wolfrum, 1:100) were used. For mouse antibodies against parvalbumin, the M.O.M. kit was used as specified by the manufacturer (Vector Labs). Tissues were washed (three times for 10–15 min each) after primary and secondary antibody (donkey anti-mouse Alexa555 (A31570) and Donkey anti-rabbit Alexa488 (A21206), 1:400, Life Technologies). incubations in 0.1% Tween-20 in PBS, and nuclei were counterstained with DAPI (1 mg/ml; D9542, Sigma-Aldrich). F-actin was labeled with rhodamine phalloidin (Life Technologies) according to the manufacturer’s instructions. For counting hair cells, specimens were dehydrated through an ethanol series, cleared with methyl salicylate:benzl benzoate (5:3) and examined by confo-cal fluorescence microscopy. For harmonin b analysis, labeled specimens were mounted and stored in Prolong Gold (Life Technologies). All samples were imaged with a Zeiss motorized system operated with LSM software (Zeiss) and equipped with 405-, 543- and 633-nm diodes along with a multiline argon laser (457 nm, 488 nm, 515 nm); an XYZ stage; and several objectives that include the Plan-NEOFLUAR 10× (NA = 0.3), Plan-NEOFLUAR 40× (NA = 1.3 oil) and Plan-APOACHROMAT 100× (NA = 1.4 oil) used. Scans were performed through a sequential (line) mode and PMT voltages dynamically regulated to compensate for signal loss due to scatter and depth limitations. Planes were captured at a resolution of 2048 × 2048 pixels and speeds of 1–2 ms per pixel. Optical volumes were deconvolved with a constrained maximum likelihood estimation algorithm and a calculated point spread function using Huygens

Professional 4.1 (Scientific Volume Imaging) running on a Mac Pro computer (Apple). Z-stack images were reconstructed and analyzed using ImageJ, Fuji and Photoshop softwares.

Statistical analyses. Data were analyzed by ANOVA with post hoc tests and Student’s t-test (SAS Institute Inc, NC or Prism 5 Graphpad Software) as noted in the figure legends. Hair cell counts were analyzed as the dependent variable separately for both inner and outer hair cells in a nested ANOVA with a two-level factorial arrangement of treatments33. The nested effect was the mice within each genotype treatment combination, the two main effect factors were cell location in the cochlea and genotype/treatment combination (combined into one variable with three levels, see Fig. 4). Adjustment for multiple comparisons conducted to separate interaction means was by a simulation method34. All data management and analysis was performed using programs and procedures in the Statistical Analysis System (SAS Institute).

29. Baker, B.F. et al. 2ʹ-O-(2-methoxy)ethyl–modified anti-intercellular adhesion molecule 1 (ICAM-1) oligonucleotides selectively increase the ICAM-1 mRNA level and inhibit formation of the ICAM-1 translation initiation complex in human umbilical vein endo-thelial cells. J. Biol. Chem. 272, 11994–12000 (1997).

30. Hastings, M.L. et al. Tetracyclines that promote SMN2 exon 7 splicing as therapeutics for spinal muscular atrophy. Sci. Transl. Med. 1, 5ra12 (2009).

31. Hardie, N.A., MacDonald, G. & Rubel, E.W. A new method for imaging and 3D recon-struction of mammalian cochlea by fluorescent confocal microscopy. Brain Res. 1000, 200–210 (2004).

32. Sage, C., Venteo, S., Jeromin, A., Roder, J. & Dechesne, C.J. Distribution of frequenin in the mouse inner ear during development, comparison with other calcium-binding proteins and synaptophysin. Hear. Res. 150, 70–82 (2000).

33. Milliken, G.A. & Johnson, D.E. Analysis of Messy Data Volume I: Designed Experiments Ch. 30, 413–423 (Lifetime Learning Publications, Belmont, California, 1984).

34. Edwards, D. & Berry, J.J. The efficiency of simulation-based multiple comparisons. Biometrics 43, 913–928 (1987).

doi:10.1038/nm.3106 nature medicine

npg

© 2

013

Nat

ure

Am

eric

a, In

c. A

ll rig

hts

rese

rved

.

Page 19: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Competing financial interests

F.R. may materially benefit financially through stock options in Isis Pharmaceuticals. M.L.H. and F.R. have patents pending with the United States Patent and Trademark Office for the ASOs and the targeting approach.

npg

© 2

013

Nat

ure

Am

eric

a, In

c. A

ll rig

hts

rese

rved

.

Page 20: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Neuron

Article

Restoration of Hearingin the VGLUT3 Knockout MouseUsing Virally Mediated Gene TherapyOmar Akil,1 Rebecca P. Seal,3 Kevin Burke,1 Chuansong Wang,4 Aurash Alemi,1 Matthew During,4 Robert H. Edwards,2

and Lawrence R. Lustig1,*1Department of Otolaryngology, Head and Neck Surgery2Department of NeurologyUniversity of California San Francisco, San Francisco, CA, 94143, USA3Department of Neurology, University of Pittsburgh, Pittsburgh, PA 15213, USA4Comprehensive Cancer Center, The Ohio State University, Columbus, OH 43210, USA

*Correspondence: [email protected]://dx.doi.org/10.1016/j.neuron.2012.05.019

SUMMARY

Mice lacking the vesicular glutamate transporter-3(VGLUT3) are congenitally deaf due to loss of gluta-mate release at the inner hair cell afferent synapse.Cochlear delivery of VGLUT3 using adeno-associ-ated virus type 1 (AAV1) leads to transgene expres-sion in only inner hair cells (IHCs), despite broaderviral uptake. Within 2 weeks of AAV1-VGLUT3delivery, auditory brainstem response (ABR) thresh-olds normalize, alongwith partial rescue of the startleresponse. Lastly, we demonstrate partial reversal ofthe morphologic changes seen within the afferentIHC ribbon synapse. These findings represent asuccessful restoration of hearing by gene replace-ment in mice, which is a significant advance towardgene therapy of human deafness.

INTRODUCTION

Hearing loss is one of the most common human sensory deficits,

with congenital hearing loss occurring in approximately 1.5 in

1,000 children (Smith et al., 2005). Of these, about half are attrib-

uted to a genetic basis (Di Domenico et al., 2011). While our

understanding of the causes of genetic hearing loss has

advanced tremendously over the past 30 years (Petersen and

Willems, 2006), treatments have advanced little over this same

time period and currently consist of hearing amplification for

mild to severe losses and cochlear implantation for severe to

profound losses (Kral and O’Donoghue, 2010). Though cochlear

implantation has profoundly influenced our treatment of children

with congenital deafness, there are still significant limitations in

function with an implant, and these results cannot compare to

native hearing (Kral and O’Donoghue, 2010). Thus, there remains

intense interest in restoring normal organ of Corti function

through techniques such as hair cell regeneration and gene

therapy (Di Domenico et al., 2011). To date, a majority of the

research in this arena has focused on cochlear hair cell regener-

ation, applicable to the most common forms of hearing loss

including presbycusis, noise damage, infection, and ototoxicity.

Several studies have now demonstrated regeneration of hair

cells in injured mice cochlea and improvement of both hearing

and balance with virally mediated delivery of Math1 (Baker

et al., 2009; Husseman and Raphael, 2009; Izumikawa et al.,

2008; Kawamoto et al., 2003; Praetorius et al., 2010; Staecker

et al., 2007). While these efforts in wild-type animals are quite

important, they still do not address the problem of an underlying

causative genetic mutation. In such a scenario, even success-

fully regenerated hair cells will still be subject to the innate

genetic mutation that led to hair cell loss in the first place. To

date, efforts to restore hearing in this type of hearing loss with

gene therapy have been met with limited success (Maeda

et al., 2009), and no study has reported the reversal of deafness

in an animal model of genetic deafness.

Previous reports have described a mouse model of hereditary

deafness, which occurs as a result of a null mutation in the

gene coding for the vesicular glutamate transporter-3 (VGLUT3)

(Obholzer et al., 2008; Ruel et al., 2008; Seal et al., 2008).

Synaptic transmission mediated by glutamate requires transport

of the excitatory amino acid into secretory vesicles by a family of

three vesicular glutamate transporters (Fremeau et al., 2004;

Takamori et al., 2002). We previously demonstrated that inner

hair cells of the cochlea express VGLUT3 and that mice lacking

this transporter are congenitally deaf (Seal et al., 2008). Hearing

loss in these mice is due to the elimination of glutamate release

by inner hair cells and hence to the loss of synaptic transmission

at the IHC-afferent nerve synapse. Subsequent studies have

shown that a missense mutation in the human gene SLC17A8,

which encodes VGLUT3, might underlie the progressive high-

frequency hearing loss seen in autosomal dominant DFNA25

(Ruel et al., 2008). Here we report the successful restoration of

hearing in the VGLUT3 knockout (KO) mouse using virally

mediated gene delivery.

RESULTS

AAV1-VGLUT3 Transfection Results in LocalizedExpression in Inner Hair CellsOur first goal was to determine the extent of transfection with the

adeno-associated virus type 1 (AAV1) within the cochlea. Using

Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc. 283

Page 21: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

an AAV1-GFP construct, there appeared to be labeling of

a variety of cell types within the cochlea, including the inner

hair cells and supporting cells using an anti-GFP antibody (Fig-

ure 1A), in a pattern similarly described by other investigators

(Jero et al., 2001; Konishi et al., 2008). Subsequently, virus con-

taining the VGLUT3 gene (AAV1-VGLUT3) was microinjected

into the cochlea using two different techniques: initially via an

apical cochleostomy (CO) and subsequently by direct injection

through the round window membrane (RWM) (Figures 1B–1E).

After delivery, RT-PCR of inner ear tissue (Figure 1C) demon-

strated strong VGLUT3 mRNA expression in the rescued whole

cochlea, organ of Corti, stria vascularis, vestibular neuroepithe-

lium, and very weakly in the spiral ganglion. Noninjected

cochleas of knockouts do not demonstrate VGLUT3 expression

as noted (Figure 1C, KO �/+RT). In contrast, under immunofluo-

rescence, inner hair cells were the only cells labeled with anti-

VGLUT3 antibody (Figure 1B).

To determine the dose dependence of VGLUT3 expression

in the IHCs, we injected either 0.6 ml or 1 ml of AAV1-VGLUT3

(2.3 3 1013 virus genomes [vg]/ml) into the cochlea (Figures 1D

and 1E). Microinjecting 1 ml of virus resulted in 100% of IHCs

labeled with anti-VGLUT3 antibody; in contrast, microinjecting

0.6 ml resulted in only �40% of IHCs labeled by the antibody.

We next sought to determine whether earlier viral delivery

would result in more robust VGLUT3 expression (Figures 1D,

1E, and 2). As shown, delivery of virus via the RWM at postnatal

day 10 (P10) results in �40% of the IHCs expressing VGLUT3

(Figures 1D, 1E, and 2), whereas similar doses (0.6 ml) of virus

injected at P1–P3 results in 100% of IHC transfected in all

animals (Figures 1D, 1E, and 2).

Delivery of AAV1-VGLUT3 Restores Normal ABR andCAP Thresholds within 14 DaysAfter verifying successful transgene expression within the IHC

without significant organ of Corti injury, we next sought to deter-

mine whether the reintroduction of VGLUT3 would lead to meas-

ureable hearing recovery (Figure 3). In these studies, only 0.6 ml

of AAV1-VGLUT3 was delivered at P10–P12. Auditory brainstem

response (ABR) thresholds were first measurable within 7 days

after viral delivery, with near normalization of thresholds

to wild-type (WT) levels within 2 weeks (P24–P26) (Figures

3A–3C). Initially a CO technique was used for viral delivery.

However, this method restored hearing in only �17% of animals

(n = 5 out of 30 animals attempted), presumably because it was

more technically challenging and due to the trauma of the

approach (see Discussion). As a result, the method was subse-

quently changed to an RWM delivery, which resulted in hearing

restoration in 100% of mice (n = 19 out of 19 mice). The time

course of hearing recovery was similar for the CO (when

successful in 17%) and the RWM delivery techniques (100% of

mice). Compound action potentials (CAPs) were also restored

within 7–14 days of viral delivery (Figure 3A). Since the loss of

VGLUT3 affects only glutamate release at the IHC synapse

(Ruel et al., 2008; Seal et al., 2008), restoration of normal ABRs

and CAPs also implies restoration of synaptic function. We

also compared the longevity of hearing recovery, defined as

the period of time between onset of hearing recovery and

when ABR thresholds become elevated >10 dB aboveWT levels,

284 Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc.

between the CO and RWMmethods (Figure 3D). In both groups,

all rescued KO mice maintained hearing for at least 7 weeks. At

28 weeks postdelivery, 40% of the mice who achieved suc-

cessful CO delivery still had hearing within 10 dB of WT mice

(n = 2/5), while only 5% of the RWM mice had the same level

of hearing (n = 1/19). Interestingly, some rescued mice in each

group, CO and RWM, maintained normal ABR thresholds up to

1.5 years. The number of animals for each rescued group at

each time point, within 10 dB of WT thresholds, is described in

the legend of Figure 3D.

We subsequently measured hearing recovery in mice injected

via the RWM at P1–P3 (Figure 3D). Due to the small size of the

cochlea, only 0.6 ml of virus could be delivered at this time point.

However, 100% of mice recovered normal ABR thresholds by

P14 (n = 19 mice). Five mice were followed for 9 months and still

maintained normal ABR thresholds at this later time point. Earlier

delivery thus not only appears to be more efficient (100% of

animals recover hearing) but also leads to greater longevity of

hearing recovery.

Bilateral AAV1-VGLUT3 Rescue Results in a LargerRecovery of Behavioral and Electrical Measuresof HearingFor an additional assay of hearing recovery, we studied the

startle response at approximately 3 weeks after viral delivery

(Figure 4). In these experiments, the AAV1-VGLUT3 delivery

was done via the RWM at age P10–P12. As expected, VGLUT3

KOmice show no startle response due to the absence of hearing.

When hearing was rescued in one ear (‘‘unilat,’’ Figure 4A), at the

loudest presentation level of 120 dB, the startle response

improved to 8% of normal, while if both ears were rescued

(‘‘bilat,’’ Figure 4A), the startle response increased to 33% of

normal, both measures being statistically different than the KO

response. Interestingly, similar amplitude growth was observed

with ABR wave I amplitudes when both ears, as opposed to

a single ear, were rescued (Figure 4B). ABR wave I latency was

also studied (Figure 4C), and while there appeared to be a trend

for reduced latency in the unilateral-rescued mice, the differ-

ences between unilateral- and bilateral-rescued and WT mice

were not significant. Thus, while ABR thresholds can be brought

to normal, ‘‘behavioral’’ thresholds and ABR amplitudes can be

improved, but not normalized, to the WT level with this rescue

technique.

AAV1-VGLUT3 Rescue Partially Restores SynapticMorphology, but Not Spiral Ganglion Cell Counts,in the Knockout MouseAs we previously demonstrated (Seal et al., 2008), at P21,

VGLUT3 KO mice show a 10%–18% decrease in spiral ganglion

(SG) neurons compared to WT mice. This decrease was still

observed in the AAV1-VGLUT3 rescued mice (RWM delivery at

P10–P12) at P21 (Figure 5A). Further, rescued mice showed no

significant differences in spiral ganglion cell size as compared

to KO mice (Figure 5B), though both were significantly less

than WT mice. To determine whether long-term hearing would

reverse this trend, we also took counts at 5 months after birth,

but again, no significant differences in SG counts or cell size

were seen in the KO versus rescued mice at this later time point

Page 22: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Figure 1. AAV1-GFP Transduction in Mice Organ of Corti

(A) AAV1-GFPwas used to assess viral delivery to the cochlea on organ of Corti surface preparations. AAV1 transfects a number of cell types including inner (IHC)

and supporting cells (white arrows, left) in the organ of Corti. The VGLUT3 KO (V3KO, middle) documents IHC labeling of transfected cells only. In the WTmouse

(right), VGLUT3 colabels IHCs (red) along AAV1-GFP-transfected cells (green).

(B) VGLUT3 expression after AAV1-VGLUT3 delivery via the RWM, delivered at P10–P12, and stained with anti-myosin 7A antibody (green, used as a hair cell

marker, left column) and anti-VGLUT3 antibody (red, middle column) and merged (right column). WTmice show IHCs labeled with both anti-Myo7A and VGLUT3

as expected (top row, right), whereas KO mice only show Myo7a expression (middle row, right). After AAV1-VGLUT3 delivery, the IHC is colabeled by both

anti-Myo7a and anti-VGLUT3 antibodies (bottom row, right) (TC, tunnel of Corti; DC, Deiter’s cells; OHC, outer hair cells; IHC, inner hair cells).

(C) RT-PCR was used to verify VGLUT3 mRNA expression in the transfected KO (first lane) and WT (second lane) mice. Rescued whole cochlear extract (R-Co)

demonstrates strong VGLUT3 mRNA expression, as does the stria vascularis + organ of Corti lane (R-SV+OC) and vestibular neuroepithelium (R-Vest). Spiral

ganglion only shows weak mRNA expression (R-SG).

(D and E) Inner hair cells labeled with anti-VGLUT3 antibody after transfection were counted, tabulating both the entire cochlea and within the base, midturn, and

apex to determine differences in regions (at P10–P12 delivery, WT n = 5, CO n = 5, RWM n = 6 and at P1–P3 delivery, RWM n = 6). At P10–P12, when 1 ml of virus

was microinjected, 100% of IHCs were labeled and similar results were seen at P1–P3 viral delivery (D) when injected with 0.6 ml. In contrast, when 0.6 ml of virus

was injected at P10–P12, approximately 40% of IHCs were labeled, with no significant differences seen between the apex, midturn, or base in the variability of

IHC labeling with the delivery technique (E).

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc. 285

Page 23: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Figure 2. VGLUT3 IHC Transfection: Early versus Late Delivery

AAV1-VGLUT3 delivery at P1–P3 versus P10–P12 are compared using similar amount of virus (0.6 ml) and examined at �P30. Anti-Myo7a antibody (green) and

anti-VGLUT3 antibody (red) are used for staining, and the merged images (yellow) are shown. As expected, IHC from WT mice show both anti-Myo7a and anti-

VGLUT3 staining (row 1), whereas KO mice only show anti-Myo7a label (row 2). Delivery of virus via the RWM at P10–P12 results in fewer IHCs expressing

VGLUT3 (row 3), whereas similar doses of virus injected at P1–P3 results in 100% of hair cells transfected in all animals (row 4) (IHC, inner hair cells).

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

(data not shown). Subsequently, spiral ganglion cell counts were

also undertaken in mice that underwent virus delivery at P1–P3.

However, despite a robust IHC transfection and early hearing

recovery (see Figures 1D, 1E, and 2), again, no differences in

SG cell counts were noted between KO and rescued mice

(data not shown). Additionally, histology (Figure 5C) documents

no obvious cochlear trauma as a result of viral delivery in the

rescued mice, as evidenced by normally appearing organ of

Corti structures with preservation of inner and outer hair cells,

supporting cells, spiral ganglion neurons (though similarly

reduced in number as nonrescuedmice), and the stria vascularis

(data not shown).

As originally reported, VGLUT3 KO mice demonstrate abnor-

mally thin, elongated ribbons in IHC synapses, though the

number of synaptic vesicles tethered to ribbons or docked at

the plasma membrane was normal (Seal et al., 2008). We thus

sought to determine whether these morphologic abnormalities

could be reversedwith hearing rescue. As shown (Figure 6, Table

1), in the rescued mice, ribbon synapses are normal in appear-

ance, taking on a more rounded shape similar to the WT, while

the nonrescued mice continue to demonstrate abnormally thin

and elongated ribbons. The rescuedmice also displayed a signif-

icantly larger number of synaptic vesicles associated with the

ribbon (19 rescued versus 14 WT, p = 0.02) (Table 1). Interest-

286 Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc.

ingly, within individual hair cells, the synaptic vesicles them-

selves demonstrated a mixture of elongated and circular mor-

phology, as opposed to all circular in the WT and all elongated

in the KO mice. However, when analyzing the average number

of docked synaptic vesicles at a ribbon synapse, rescued

animals did not show a significant difference between the WT

and KO mice (Table 1). While these results demonstrate only

a partial reversal of the synaptic changes seen in the KO mouse

ribbon synapse, it is enough to recover ABR thresholds to theWT

levels in the rescued KO mice.

DISCUSSION

These studies document the successful rescue of the deafness

phenotype in a mouse model of inherited deafness. With viral

delivery of VGLUT3 at P10–P12 in the KO mouse, ABR thresh-

olds normalize within 7–14 days and remain in this range for at

least 7 weeks, with two mice maintaining auditory thresholds

for as long as a year and a half in this current study. Earlier

delivery, at P1–P3, results in an even more robust IHC transfec-

tion and long-lived hearing recovery in this mouse model.

One unexpected result from these investigations was the

differential finding of more widespread expression of green fluo-

rescent protein (GFP) after AAV1-GFP transfection as compared

Page 24: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Figure 3. Hearing Restoration in the VGLUT3 KO Mice

(A) Representative ABR and CAP tracings fromWT, KO, and rescued KO mice after delivery of 0.6 ml AAV1-VGLUT3. Waveforms from the WT and rescued mice

appear similar, while KO mice show no ABR and no CAP responses. ‘‘I’’ indicates the location of ABR wave I.

(B) Rescued mice begin to show hearing recovery within 7 days postdelivery (P17–P19), with near normalization of ABR thresholds by 14 days postdelivery

(P24–P26). Hearing recovery is seen with both CO and RWM delivery through 69 weeks. Black arrows indicate absence of ABR threshold of the KO above 92 dB

(the maximum level that our recording equipment can measure).

(C) At 40 days postdelivery (P50–P52), similar levels of recovery are noted at 8 and 16 kHz, while at 32 kHz, CO delivery appears to result in slightly elevated

thresholds, though still significantly better than KO.

(D) Hearing longevity, as defined as the number of mice with ABR threshold levels within 10 dB of WT levels at each time point, is measured for RWM at P10–P12,

CO at P10–P12, and RWM at P1–P3. For viral delivery at P10–P12, CO delivery results in longer-lasting hearing recovery on average than RWM delivery, with

some rescued KO in each group maintaining ABR thresholds to within 10 dB of WT animals beyond 28 weeks. However, 100% of rescued KO mice with RWM

delivery recover hearing, while only 17% of CO-rescued mice recover hearing. The number of animals with viral delivery at P10–P12 at each time point is the

following: RWM 0–9 weeks n = 19, 10–13 weeks n = 12, 14–24 weeks n = 8, 24–28 weeks n = 1, 28+ weeks n = 1; CO 0–7 weeks n = 5, 8–14 weeks n = 4,

15–20 weeks n = 3, 21–28 weeks n = 2, 28+ weeks n = 1. In contrast, in mice undergoing viral delivery at P1–P3, all animals maintain ABR threshold recovery at

least through 9 months (n = 5).

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

to isolated VGLUT3 expression restricted to the IHC after

AAV1-VGLUT3 transfection (Figures 1A–1C and 2). RT-PCR

results demonstrate that after AAV1-VGLUT3 delivery, there is

also more widespread VGLUT3 mRNA transcription than in just

IHCs (Figure 1C). These results suggest that there is a posttran-

scriptional regulatory mechanism acting on VGLUT3 mRNA,

which leads to selective expression of the protein only within

IHCs. Several types of posttranscriptional regulation have been

described within the cochlea, and whether this specific mecha-

nism involves microRNA inactivation (Elkan-Miller et al., 2011),

transcription factor regulation (Masuda et al., 2011), or another

process remains to be determined. Such a mechanism, if appro-

priately elucidated and exploited, could theoretically allow the

expression (or conversely suppression) of a number of different

proteins within the inner ear to alter function in pursuit of hearing

preservation.

Another interesting finding was the variable success with the

CO as compared to the RWM delivery technique. As noted, we

initially started with an apical CO delivery method but aban-

doned it due to the low success rate of hearing restoration

(17% of animals). Subsequently, we changed to an RWM

delivery technique for several reasons; this would be the most

likely method of delivery in any future human studies, and it

was less likely to be traumatic, as evidenced by a number of

recent human studies looking at hearing preservation with round

window insertion of cochlear implants (von Ilberg et al., 2011). In

Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc. 287

Page 25: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Figure 4. Behavioral Measures and Physiologic Growth after AAV1-VGLUT3 Rescue

(A) Startle responses were studied as an additional behavioral measure of hearing recovery 3 weeks after viral delivery (P31–P33). While none of the rescuedmice

recover startle responses to WT levels, they nonetheless develop a response, which increases when both ears are rescued; the louder the sound delivered (100,

110, and 120 dB presentation levels), themore robust the startle response was, with rescue of both ears (bilat) creating a larger response than a single ear (unilat).

These differences were statistically significant.

(B) Similar growth was seen for ABR wave I amplitudes when both ears (bilat) as compared to a single ear (unilat) were rescued, though again not as robust as in

the WT mice.

(C) ABRwave I latency wasmeasured, which showed no statistically significant differences between unilateral, bilateral, rescued, andWTmice, though there was

a trend toward an increased latency in the rescued compared to WT mice.

(D) Representative ABR tracings documenting increasing wave I amplitudes from the KO, unilaterally rescued KO (unilat), bilaterally rescued KO (bilat), and WT

mice.

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

fact, the change in technique resulted in hearing restoration in

100% of animals attempted. We believe the likely difference in

success between the two techniques relates to the degree of

trauma induced by each method. With a cochleostomy, a sepa-

rate hole into the scala through bone must be created, which by

its nature is traumatic, despite our best efforts to minimize

trauma. In contrast, an RWM injection simply involves piercing

the membrane and sealing it with fascia after viral delivery.

However, we were histologically unable to see any obvious

differences between the ears of animals with andwithout hearing

rescue in the cochleostomy group (data not shown) and there

may be other reasons for the variable success between the

two techniques. Further, we noted that even earlier delivery via

the RWM at P1–P3, as opposed to P10–P12, resulted in hearing

recovery that was more consistently long lived, with all mice

288 Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc.

followed out through 9 months showing ongoing normal ABR

thresholds (Figure 3D).

Transgene expression with AAV1 should theoretically last for

a year or longer (Henckaerts and Linden, 2010). However, it is

not entirely clear why there is a variable loss of hearing after

7 weeks, regardless of delivery technique at the later P10–P12

delivery time point (Figure 3D). We analyzed individual cochleae

but did not see histologic evidence of active inflammation in

those animals that lost hearing and IHCs still expressing VGLUT3

protein. Further, spiral ganglion counts did not significantly differ

in animals with and without hearing. One possibility could be due

to the trauma of viral delivery, with gradual reopening of the

delivery site (RWM or cochleostomy) leading to a perilymphatic

leak with resulting hearing loss. Such a lesion might not be

detectable on histology. Another possible explanation may be

Page 26: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Figure 5. Spiral Ganglion Cell Counts

In these studies, the AAV1-VGLUT3 delivery was done at P10–P12 and at P1–P3. Spiral ganglion (SG) cell counts in rescued mice versus KO and WT mice were

undertaken at P21 (A and B) and at 5 months after birth (data not shown) for P10–P12 virus delivery and at twomonths after birth (data not shown) for P1–P3 virus

delivery. AAV1-VGLUT3 transfection did not result in a reversal of the SG neuronal loss in either age group. No significant differences were seen between the

groups. These results also document the lack of trauma (C) associated with virally mediated delivery of VGLUT3 through the RWM, with normal organ of Corti

morphology, including inner and outer hair cells, spiral ganglion neurons, and stria vascularis (data not shown).

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

due to transgene inactivation, by a hypothetical mechanism such

as microRNA inactivation or methylation. Clearly, if one hopes to

consistently achieve long-term transgene expression within the

ear, which will be critical for application of this technique in

humans, this variable will need to be better understood and

controlled, particularly at later ages of delivery.

It is interesting to note that the lower dose of virus used for

most of the studies performed (0.6 ml), delivered at P10–P12,

caused VGLUT3 expression in only �40% of IHCs (Figures 1D

and 1E), and yet this was enough to restore ABR thresholds to

WT levels for click responses and to near normal for pure tone

thresholds (Figures 3A–3C). Similar results have been docu-

mented in other models of hearing recovery after noise exposure

(Kujawa and Liberman, 2009; Lin et al., 2011), in which even

‘‘reversible’’ noise exposure with recovery of auditory thresholds

leads to long-term afferent nerve terminal degeneration while

retaining ‘‘normal’’ auditory thresholds. Similar findings with re-

gard to the discrepancy of ABR threshold and amplitudes have

also been shown from mutant mice lacking synaptic ribbons

(Buran et al., 2010). However, correlative studies in human

temporal bones suggest that cochlear implants in humans can

still function very effectively despite significant spiral ganglion

neuron loss, allowing for meaningful speech and sound trans-

mission (Gassner et al., 2005; Khan et al., 2005). Thus, complete

normalization of all cellular abnormalities may ultimately not be

required for the technique to be successful in humans, though

this should remain a goal for animal studies going forward.

The KO mice develop an unusual appearing ribbon that is thin

and elongated, as noted here and previously (Seal et al., 2008). A

similar ribbon morphologic pattern, flat and plate-like, is seen in

the Otoferlin KO mouse (Roux et al., 2006). As Otoferlin is also

critical in glutamate release at the IHC synapse, this implies

that lack of physiologic activity of the synapse results in such

a flat ribbon appearance. In the rescued mice, while the ribbon

itself appeared normal, we did still see a mixture of elongated

and circular vesicles within the transfected IHCs, as opposed

to all circular in theWT and all elongated in the KOmice, implying

that there may still be differences in transmitter release in the

rescued versus WT mice.

Another interesting finding with regard to the ribbon synapse

was the larger numbers of synaptic vesicles that were associ-

ated with the ribbon seen in the rescued mice, as well as the

mixture of elongated and circular vesicles observed. The data

shows that much larger quantities of VGLUT3 mRNA are being

produced in the rescued as compared to theWTmice (Figure 1C,

RT-PCR data) and suggests, though does not prove, an associ-

ation between increased mRNA levels and vesicle number. We

believe that circular vesicles represent properly packaged vesi-

cles, while the elongated vesicles are improperly packaged

vesicles. Perhaps continuous production of VGLUT3 by the

Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc. 289

Page 27: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Figure 6. Partial Normalization of Synaptic Ribbon Morphology in the Rescued Mice

Electron microscopy was performed on broken serial thin sections of the synaptic region of the IHCs, which were cut in a horizontal plane parallel to the basilar

membrane in the same orientation for the WT, KO, and rescued KO mice. As previously reported (Seal et al., 2008), VGLUT3 KO mice demonstrate abnormally

thin, elongated ribbons in IHC synapses (middle column), as compared to WT littermates (left column). In the rescued mice (right column, single ribbon shown on

top, double on the bottom), ribbon synapses are normal appearing, taking on amore rounded shape similar to theWT. Black arrows point to the normal or circular

synaptic vesicles and white arrows point to elongated synaptic vesicles.

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

constitutive CBA promoter driving transfected VGLUT3 pro-

duction prevents the IHC from properly packaging the vesicles

at a normal rate, leading to a higher number as well as a mixture

of regular and irregular-appearing vesicles. Another possibility

is that the incomplete transfection rate of IHCs (40% of IHCs

labeled at the doses used for these morphology studies) led to

the heterogeneity of the ribbon morphology seen.

The observed growth on behavioral and electrical measures

seen with bilateral, as opposed to unilateral, rescue (RWM

delivery at P10–P12; Figure 4) was an unexpected finding. While

none of the animals had complete normalization of ABR ampli-

tude and startle-response levels, the amplitude growth does

imply that bilateral input increases the auditory response cen-

trally. An analogous phenomenon is seen with ‘‘binaural summa-

tion’’ and clinically in patients who wear two hearing aids as

opposed to one and report lower levels of amplification required

(Noble, 2010; Steven Colburn et al., 2006) and suggests that the

response seen in these studies is physiologic. Recent studies

have localized VGLUT3 to various structures in the brainstem,

including cochlear nucleus (Fyk-Kolodziej et al., 2011) as well

as the LSO and MNTB (Lee et al., 2011). It is certainly possible

that deficits within auditory brainstem signal pathways could

be contributing to the inability to restore the startle response to

WT levels.

The failure of the technique to reverse the spiral ganglion cell

loss seen in the VGLUT3 KO mice when delivered at P10–P12

is not surprising (Figure 5), given that hair cell activity and afferent

290 Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc.

stimulation can provide a trophic effect on SG survival. This is

probably at least partly due to the fact that virus was delivered

at �P10 with subsequent ABR threshold recovery at �P17–

P24, after spiral ganglion neuronal degradation has begun

(Seal et al., 2008). This also implies that in order for SG neurons

to be preserved at normal levels, intervention would probably

have to occur earlier. Further, with only �40% of IHCs express-

ing VGLUT3 (using the lower concentration of virus, delivered at

P10–P12), there are still many spiral ganglion neurons not

receiving afferent input, which also probably impacts this result

as well. Wewere thus surprised that even earlier delivery of virus,

at P1–P3, which resulted in relatively early onset of hearing,

measureable by P14, with 100% of IHC expressing VGLUT3,

also did not lead to restoration of SG cell counts to WT levels.

Perhaps there are in utero factors that also help maintain SG

numbers, or even a small delay in hearing onset can lead to

SG loss.

Lastly, while the mutant mouse in the current study and the

hearing loss described in patients with DFNA25 are both due

to mutations in the gene coding for VGLUT3, the comparison

may not be straightforward. First, it is not certain that the

missense mutation described in SLC17A8 is the cause of the

hearing loss seen in DFNA25, though a strong correlation was

observed (Ruel et al., 2008). Second, the null mutation studied

in these experiments would represent a much more severe

phenotype than the missense mutation described as potentially

causative for DNFA25. Thus, whether this technique could

Page 28: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Table 1. Transmission Electron Micrographic Analysis of the

Ribbon Synapse

WT KO Rescued KO

Number of inner hair

cells examined

58 50 61

Total number of

ribbons

18 20 18

Number of floating

ribbons

3 3 4

Number of normal

ribbons

18 0 17

Number of elongated

ribbons

0 20 1

Average number of

synaptic vesicles

associated with the

ribbon

14.44 ± 1.92 16.26 ± 2.85 19.17 ± 4.52

Shape of the synaptic

vesicles

All circular All elongated Mixture of

circular +

elongated

Average number of

docked synaptic

vesicles

2.28 ± 1.27 1.56 ± 0.92 2 ± 0.84

This analysis compares the ribbon synapse and vesicles inWT versus KO

versus AAV1-VGLUT3-rescued KO mice. Electron microscopy was per-

formed on broken serial thin sections of the synaptic region of the

IHCs, which were cut in a horizontal plane parallel to the basilar

membrane with identical handling and orientation for the WT, KO, and

rescued KO. When differentiating vesicles as either ‘‘circular’’ or ‘‘elon-

gated,’’ a vesicle was defined as ‘‘circular’’ when the perpendicular diam-

eters were found to bewithin 50%of each other. Any vesicle with unequal

perpendicular dimensions (>50% difference) was counted as ‘‘elon-

gated,’’ and any ribbon synaptic body that had a length greater than three

times its greatest width was considered to be ‘‘abnormal’’ in appearance.

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

ultimately be beneficial to patients with DFNA25 remains

unclear. Despite these differences, as our study documents

restoration of normal ABR levels in such a null mutant model, it

nonetheless represents an important initial step for the potential

treatment of inherited deafness.

EXPERIMENTAL PROCEDURES

Animals

VGLUT3 null mutant mice were generated as described in a C57 (Seal et al.,

2008) strain then backcrossed with FVB mice (less than seven generations)

to obtain a homogeneous genetic background. P1–P12 mice were used for

AAV1-VGLUT3 delivery. All procedures and animal handling complied with

NIH ethics guidelines and approved protocol requirements at the University

of California, San Francisco (IACUC).

All surgical procedures were done in a clean, dedicated space. Instruments

were thoroughly cleaned with 70% ETOH and autoclaved prior surgery.

Surgery was carried out under a Leica MZ95 dissecting scope and animals

were situated with neck extended over solid support. Mice were anesthetized

by intraperitoneal injection of a mixture of ketamine hydrochloride (Ketaset,

100 mg/kg), xylazine hydrochloride (Xyla-ject, 10 mg/kg), and acepromazine

(2 mg/kg) and boosted with one-fifth the original dose as required. Depth

of anesthesia was continuously checked by deep tissue response to toe

pinch. Body temperature was maintained with a heating pad and monitored

with a rectal probe throughout procedures. Preoperatively and every 24 hr

postoperatively, animals were given subcutaneous carprofen analgesia

(2 mg/kg) to manage inflammation and pain. Animals were closely monitored

for signs of distress and abnormal weight loss postoperatively.

AAV1-VGLUT3 Viral Construct

Mouse VGLUT3 cDNA was subcloned into the multiple cloning site of vector

AM/CBA-WPRE-BGH (kindly provided by R. Palmiter). Human embryonic

kidney 293 cells were cotransfected with three plasmids—AAV-mVGLUT3

plasmid, appropriate helper plasmid-encoding rep and AAV1 cap genes,

and adenoviral helper pF D6—using standard CaPO4 transfection. Cells

were harvested 60 hr after transfection, cell pellets were lysed with sodium

deoxycholate, and AAV vectors were purified from the cell lysate by ultracen-

trifugation through an iodixanol density gradient, then concentrated and

dialyzed against phosphate-buffered saline (PBS), as previously described

(Cao et al., 2009, 2010; Lawlor et al., 2009). Vectors were titered using real-

time PCR (ABI Prism 7700; Applied Biosystems), and purity of vector stocks

was confirmed by running a 10 ml sample on sodium dodecyl sulfate polyacryl-

amide gel electrophoresis and staining with Coomassie blue.

Surgical Procedures

Round Window Membrane Injection

Animalswere anesthetized, the left ear was approached via a dorsal incision as

described by Duan et al. (2004). A small hole was made in the bulla with an

18G needle and expanded as necessary with forceps and the round window

membrane (RWM) was identified. The RWM was gently punctured with a

borosilicate capillary pipette and remained in place until efflux stabilized. A

fixed volume of AAV1-VGLUT3 (0.6 ml or 1.0 ml of a 2.33 1013 vg/ml) previously

drawn into the fine pipette was gently injected through RWM over 1–2 min.

After pulling out the pipette, the RWM niche was quickly sealed with fascia

and adipose tissue. The bulla was sealed with dental cement (Dentemp,

Majestic Drug Company) and the wound was sutured in layers with a 5-0

absorbable chromic suture (Ethicon).

Cochleostomy Delivery

The right ear was approached via ventral, paramedian incision in the neck as

described by Jero et al. (2001). The injection method was similar to the RWM

except that the hole in bulla was made slightly more anterior and larger, to

directly approach the space above the stapedial artery. Injection of virus

was made into the apical turn. We used a 0.5 mm drill bit to gently thin the

bone of the otic capsule where the stria vascularis could be slightly visualized

as a brownish stripe. Once enough bonewas shaved and a slight fluid interface

became visible, 0.6 ml VGLUT3-AAV1 (2.3 3 1013 vg/ml) was pipetted into the

hole over a period of 1–2 min. After application, the hole in the cochlea was

sealed with a small amount of bone wax. After it dried, a small amount of sterile

tissue glue was applied to the bone wax and the bulla was sealed and the

wound was sutured as described above.

Auditory Testing

Acoustic Startle Response Testing

Acoustic startle responses of VGLUT3KO (n = 5),WT littermate (n = 5), rescued

VGLUT3 KO unilateral (n = 5), and rescued VGLUT3 KO bilateral (n = 5) mice

were measured as previously described (Seal et al., 2008). In brief, in darkened

startle chambers (SR-LAB hardware and software, San Diego Instruments),

piezoelectric sensors located under the chambers detect and measure the

peak startle response. Mice were acclimatized to the startle chambers by

presentation of a 70 dB white noise for 5min and then exposed to sound inten-

sities of 100 dB, 110 dB, and 120 dB (each with a 0ms rise time, 40ms plateau,

and 0ms fall time), presented in pseudorandom order with intersound intervals

of 10–50 s. Each run was repeated eight times. Average peak startle amplitude

at each sound level was calculated from eight runs. Final results were calcu-

lated as a percentage of WT mice at the 120 dB presentation level. Statistical

significance between measures was determined using a Student’s t test with

significance defined as p < 0.05.

Auditory Brainstem Response Recording

Sounds were presented and ABRs were tested in free-field conditions as

previously described in a sound-proofed chamber (Akil et al., 2006; Fremeau

et al., 2004). ABR thresholds were determined postoperatively at varying

time points, as early as 4 days after viral delivery for P10–P12 mice. The

mean value of thresholds checked by visual inspection and computer analysis

Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc. 291

Page 29: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

was defined as ABR hearing threshold for click and 8, 16, and 32 kHz tone

stimuli.

Compound Action Potential Recording

For the CAP recording, a ventral surgical approach (Jero et al., 2001) was used

to expose the right cochlea 7–14 days after AAV1-VGLUT3 delivery to the inner

ear of the P10–P12mice, including KO (n = 5), rescued KO (n = 8), andWT litter-

mates (n = 5). A fine Teflon-coated silver wire recording electrode was placed

in the round window niche, and the ground electrode was placed in the soft

tissue of the neck. The sound stimulus was generated with Tucker-Davis

System II hardware and software (Tucker-Davis Technologies).

Immunofluorescence

Immunofluorescence studies were conducted similarly for whole-mount and

cochlear sections with the following differences.

Cochlear Whole Mount

Mice cochleaewere perfusedwith 4%PFA in 0.1MPBS (pH7.4) and incubated

in the fixative for 2 hr at 4�C. The cochleae were subsequently rinsed with PBS

three times for 10min and thendecalcifiedwith 5%EDTA in 0.1MPBS. Theotic

capsule, the lateral wall, tectorial membrane, and Reissner’s membrane were

removed in that order. The remaining organ of Corti was further dissected into

a surface preparation (microdissected into individual turns), then preincubated

for 1 hr in PBS containing 0.25% Triton X-100 and 5% normal goat serum

(blocking buffer). The whole mount was then incubated with rabbit anti-myosin

VIIa antibody (a hair cell-specificmarker) (Proteus Biosciences Cat 25-6790) at

a dilution of 1:50 in blocking buffer and guinea pig anti-VGLUT3 antibody (a gift

from Dr. Robert Edward, Department of Neurology, UCSF) at 1:5,000. After an

overnight incubation at 4�C, the cochlear whole mount was rinsed twice for

10 min with PBS and then incubated for 2 hr in goat anti-rabbit IgG conjugated

to Cy2 and goat anti-guinea pig IgG conjugated to Cy3 diluted to 1:4,000 in

PBS. Specimens were next rinsed in PBS twice for 10 min and mounted on

glass slides in a mounting solution containing DAPI (nucleus stain) and

observed under an Olympus microscope with confocal immunofluorescence.

For inner hair cell counts, the cochlear whole mounts were visualized under

a microscope equipped with epifluorescence, using a 403 objective. To quan-

tify the number of IHC transfected with AAV1-VGLUT3, we labeled specimens

with anti-VGLUT3 antibody, and IHCs were manually counted in the cochlear

whole mount and in the base, midturn, and apex. For GFP labeling, surface

preparation (cochlea whole mount) was incubated with a rabbit anti-GFP anti-

body (Invitrogen A11122) at 1:250. After an overnight incubation at 4�C, thecochlea sections were rinsed twice for 10 min with PBS and then incubated

for 2 hr in goat anti-rabbit IgG conjugated to Cy2 diluted to 1:4,000 in PBS.

They were then rinsed in PBS twice for 10 min and mounted on glass slides

in amounting solution containingDAPI and observed under anOlympusmicro-

scope with confocal immunofluorescence.

Cochlear Sections

Mice were anesthetized and their cochleae were isolated, dissected, perfused

through oval and round windows by 2% paraformaldehyde in 0.1 M PB at

pH 7.4, and incubated in the same fixative for 2 hr. After fixation, the cochleae

were rinsed with PBS and immersed in 5% EDTA in 0.1 M PB for decalcifica-

tion. When the cochleae were completely decalcified, they were incubated

overnight in 30% sucrose for cryoprotection. The cochleae then were

embedded in OCT Tissue Tek Compound (Miles Scientific). Tissues were

cryosectioned at 10–12 mm thickness, mounted on Superfrost microscope

slides (Erie Scientific), and stored at �20�C until use. Sections were then

double labeled as described above (see cochlear whole mount). Slides were

then mounted in a 1:1 mixture of PBS and glycerol before being coverslipped.

Slides treated with the same technique but without incubation with the primary

antibody used as a control.

Histology

Light Microscopy

Cochleae were isolated from deeply anesthetized WT, VGLUT3 KO, and

rescued KO mice, perfused through oval and round windows with 2.5% para-

formaldehyde and 1.5% glutaraldehyde in 0.1 M PB at pH 7.4, and incubated

overnight at 4�C with slow agitation in fixative. The cochleae were rinsed with

0.1 M PB and postfixed in 1% osmium tetroxide and 1.5% potassium ferricy-

anide (for improved contrast) for 2 hr. The cochleae subsequently were

292 Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc.

immersed in 5% EDTA (0.2 M). The decalcified cochlea were dehydrated in

ethanol and propylene oxide and embedded in Araldite 502 resin (Electron

Microscopy Sciences) and sectioned at 5 mm. After sections were stained

with toluidine blue, they were mounted in Permount (Fisher Scientific) on

microscope slides.

Electron Microscopy

Electron microscopy was performed as previously described (Akil et al., 2006)

on broken serial thin sections of the synaptic region of the IHCs, which were

cut in a horizontal plane parallel to the basilar membrane. In this study, the

cochleae were all handled and cut exactly the same and the same protocol

and orientation for the WT, KO, and rescued KOwere applied when examining

and visualizing the synaptic ribbons and vesicles. The morphological assess-

ment of ribbons and vesicles was performed as described by Roux et al. (2006)

using 50–61 IHCs and 17–20 different IHC ribbon synapses from three WT,

three KO, and three rescued KO mice. Sections were stained with uranyl

acetate and lead citrate and examined under 60kV in a JEOL-JEM 100S trans-

mission electron microscope. The number of vesicles tethered to the ribbon

included all the vesicles within 30 nm of the ribbon. All the vesicles clearly

located immediately below the ribbon were considered to be docked in our

two-dimensional (2D) estimation.

Spiral Ganglion Cell Counts

Spiral ganglion cell numerical density assessmentwas calculated as described

by Leake et al. (2011) to accurately estimate the number of nuclei in a given

volumeof tissue. For this analysis, threesets of three serial sections (5mmthick-

ness)were collected from the base,midturn, and the apex of fourWT, threeKO,

and four rescued KO cochlea. Adjacent serial sections were compared, and

new nuclei of spiral ganglion neurons that appear in the second section were

counted. Statistical differences were measured using a Student’s t test.

RT-PCR of the Cochlea

Cochlea from WT, VGLUT3 KO, and rescued KO were dissected. The total

RNA was extracted from the whole cochlea, organ of Coti + stria vascularis,

spiral ganglion, and vestibular epithelium (Trizol, Invitrogen) and reverse tran-

scribed with superscript II RNase H� (Invitrogen) for 50 min at 42�C, usingoligodT primers (Akil et al., 2006). Reactions without the reverse transcriptase

enzyme (�RT) were performed as control. Two microliters of RT reaction

product were used for subsequent polymerase chain reaction (PCR; Taq

DNA Polymerase, Invitrogen) of 35 cycles using the following parameters:

94�C for 30 s, 60�C for 45 s, 72�C for 1 min, followed by a final extension of

72�C for 10min and storage at 4�C. Primers were designed to amplify a unique

sequence of VGLUT3 isoform of 759 bp. The PCR primers that were used for

mouse include VGLUT3 (GenBank accession number AF510321.1: forward-

[gctggaccttctatttgctctta] and reverse- [tctggtaggataatggctcctc]). Analysis of

each PCR sample was then performed on 2% agarose gels containing

0.5 mg/ml ethidium bromide. Gels were visualized using a digital camera and

image processing system (Kodak). Candidate bands were cut out and the

DNA was extracted (Qiaquick gel extraction kit, QIAGEN) and sequenced

(Elim Biopharmaceuticals). The PCR product was then compared directly to

the full VGLUT3 sequence for identity.

ACKNOWLEDGMENTS

We thank Dr. Diana Bautista and Dr. Makoto Tsunozaki (UC Berkeley) for

critical advice and the use of their startle response chamber. The authorswould

like to acknowledge the financial support provided by Hearing Research.

Accepted: May 1, 2012

Published: July 25, 2012

REFERENCES

Akil, O., Chang, J., Hiel, H., Kong, J.H., Yi, E., Glowatzki, E., and Lustig, L.R.

(2006). Progressive deafness and altered cochlear innervation in knock-out

mice lacking prosaposin. J. Neurosci. 26, 13076–13088.

Baker, K., Brough, D.E., and Staecker, H. (2009). Repair of the vestibular

system via adenovector delivery of Atoh1: a potential treatment for balance

disorders. Adv. Otorhinolaryngol. 66, 52–63.

Page 30: UC SF · Auditory hair cell replacement and hearing improvement by Atoh1 gene therapy in deaf mammals Izumikawa et al. (2005) Auditory hair cell replacement and hearing improvement

Neuron

Hearing Restoration in the VGLUT3 KO Mouse

Buran, B.N., Strenzke, N., Neef, A., Gundelfinger, E.D., Moser, T., and

Liberman, M.C. (2010). Onset coding is degraded in auditory nerve fibers

from mutant mice lacking synaptic ribbons. J. Neurosci. 30, 7587–7597.

Cao, L., Lin, E.J., Cahill, M.C., Wang, C., Liu, X., and During, M.J. (2009).

Molecular therapy of obesity and diabetes by a physiological autoregulatory

approach. Nat. Med. 15, 447–454.

Cao, L., Liu, X., Lin, E.J., Wang, C., Choi, E.Y., Riban, V., Lin, B., and During,

M.J. (2010). Environmental and genetic activation of a brain-adipocyte

BDNF/leptin axis causes cancer remission and inhibition. Cell 142, 52–64.

Di Domenico, M., Ricciardi, C., Martone, T., Mazzarella, N., Cassandro, C.,

Chiarella, G., D’Angelo, L., and Cassandro, E. (2011). Towards gene therapy

for deafness. J. Cell. Physiol. 226, 2494–2499.

Duan, M., Venail, F., Spencer, N., and Mezzina, M. (2004). Treatment of

peripheral sensorineural hearing loss: gene therapy. Gene Ther. 11 (Suppl

1 ), S51–S56.

Elkan-Miller, T., Ulitsky, I., Hertzano, R., Rudnicki, A., Dror, A.A., Lenz, D.R.,

Elkon, R., Irmler, M., Beckers, J., Shamir, R., and Avraham, K.B. (2011).

Integration of transcriptomics, proteomics, and microRNA analyses reveals

novel microRNA regulation of targets in the mammalian inner ear. PLoS ONE

6, e18195.

Fremeau, R.T., Jr., Kam, K., Qureshi, T., Johnson, J., Copenhagen, D.R.,

Storm-Mathisen, J., Chaudhry, F.A., Nicoll, R.A., and Edwards, R.H. (2004).

Vesicular glutamate transporters 1 and 2 target to functionally distinct synaptic

release sites. Science 304, 1815–1819.

Fyk-Kolodziej, B., Shimano, T., Gong, T.W., and Holt, A.G. (2011). Vesicular

glutamate transporters: spatio-temporal plasticity following hearing loss.

Neuroscience 178, 218–239.

Gassner, H.G., Shallop, J.K., and Driscoll, C.L. (2005). Long-term clinical

course and temporal bone histology after cochlear implantation. Cochlear

Implants Int. 6, 67–76.

Henckaerts, E., and Linden, R.M. (2010). Adeno-associated virus: a key to the

human genome? Future Virol 5, 555–574.

Husseman, J., and Raphael, Y. (2009). Gene therapy in the inner ear using

adenovirus vectors. Adv. Otorhinolaryngol. 66, 37–51.

Izumikawa, M., Batts, S.A., Miyazawa, T., Swiderski, D.L., and Raphael, Y.

(2008). Response of the flat cochlear epithelium to forced expression of

Atoh1. Hear. Res. 240, 52–56.

Jero, J., Mhatre, A.N., Tseng, C.J., Stern, R.E., Coling, D.E., Goldstein, J.A.,

Hong, K., Zheng, W.W., Hoque, A.T., and Lalwani, A.K. (2001). Cochlear

gene delivery through an intact round window membrane in mouse. Hum.

Gene Ther. 12, 539–548.

Kawamoto, K., Ishimoto, S., Minoda, R., Brough, D.E., and Raphael, Y. (2003).

Math1 gene transfer generates new cochlear hair cells in mature guinea pigs

in vivo. J. Neurosci. 23, 4395–4400.

Khan, A.M., Handzel, O., Burgess, B.J., Damian, D., Eddington, D.K., and

Nadol, J.B., Jr. (2005). Is word recognition correlated with the number of

surviving spiral ganglion cells and electrode insertion depth in human subjects

with cochlear implants? Laryngoscope 115, 672–677.

Konishi, M., Kawamoto, K., Izumikawa, M., Kuriyama, H., and Yamashita, T.

(2008). Gene transfer into guinea pig cochlea using adeno-associated virus

vectors. J. Gene Med. 10, 610–618.

Kral, A., and O’Donoghue, G.M. (2010). Profound deafness in childhood.

N. Engl. J. Med. 363, 1438–1450.

Kujawa, S.G., and Liberman, M.C. (2009). Adding insult to injury: cochlear

nerve degeneration after ‘‘temporary’’ noise-induced hearing loss.

J. Neurosci. 29, 14077–14085.

Lawlor, P.A., Bland, R.J., Mouravlev, A., Young, D., and During, M.J. (2009).

Efficient gene delivery and selective transduction of glial cells in the mamma-

lian brain by AAV serotypes isolated from nonhuman primates. Mol. Ther. 17,

1692–1702.

Leake, P.A., Hradek, G.T., Hetherington, A.M., and Stakhovskaya, O. (2011).

Brain-derived neurotrophic factor promotes cochlear spiral ganglion cell

survival and function in deafened, developing cats. J. Comp. Neurol. 519,

1526–1545.

Lee, J.H., Pradhan, J., Maskey, D., Park, K.S., Hong, S.H., Suh, M.W., Kim,

M.J., and Ahn, S.C. (2011). Glutamate co-transmission fromdevelopingmedial

nucleus of the trapezoid body—lateral superior olive synapses is cochlear

dependent in kanamycin-treated rats. Biochem. Biophys. Res. Commun.

405, 162–167.

Lin, H.W., Furman, A.C., Kujawa, S.G., and Liberman, M.C. (2011). Primary

neural degeneration in the guinea pig cochlea after reversible noise-induced

threshold shift. J. Assoc. Res. Otolaryngol. 12, 605–616.

Maeda, Y., Sheffield, A.M., and Smith, R.J. (2009). Therapeutic regulation of

gene expression in the inner ear using RNA interference. Adv.

Otorhinolaryngol. 66, 13–36.

Masuda, M., Dulon, D., Pak, K., Mullen, L.M., Li, Y., Erkman, L., and Ryan, A.F.

(2011). Regulation of POU4F3 gene expression in hair cells by 50 DNA in mice.

J. Neurosci. 197, 48–64.

Noble, W. (2010). Assessing binaural hearing: results using the speech, spatial

and qualities of hearing scale. J. Am. Acad. Audiol. 21, 568–574.

Obholzer, N., Wolfson, S., Trapani, J.G., Mo, W., Nechiporuk, A., Busch-

Nentwich, E., Seiler, C., Sidi, S., Sollner, C., Duncan, R.N., et al. (2008).

Vesicular glutamate transporter 3 is required for synaptic transmission in

zebrafish hair cells. J. Neurosci. 28, 2110–2118.

Petersen, M.B., and Willems, P.J. (2006). Non-syndromic, autosomal-

recessive deafness. Clin. Genet. 69, 371–392.

Praetorius, M., Hsu, C., Baker, K., Brough, D.E., Plinkert, P., and Staecker, H.

(2010). Adenovector-mediated hair cell regeneration is affected by promoter

type. Acta Otolaryngol. 130, 215–222.

Roux, I., Safieddine, S., Nouvian, R., Grati, M., Simmler, M.C., Bahloul, A.,

Perfettini, I., Le Gall, M., Rostaing, P., Hamard, G., et al. (2006). Otoferlin,

defective in a human deafness form, is essential for exocytosis at the auditory

ribbon synapse. Cell 127, 277–289.

Ruel, J., Emery, S., Nouvian, R., Bersot, T., Amilhon, B., Van Rybroek, J.M.,

Rebillard, G., Lenoir, M., Eybalin, M., Delprat, B., et al. (2008). Impairment of

SLC17A8 encoding vesicular glutamate transporter-3, VGLUT3, underlies

nonsyndromic deafness DFNA25 and inner hair cell dysfunction in null mice.

Am. J. Hum. Genet. 83, 278–292.

Seal, R.P., Akil, O., Yi, E., Weber, C.M., Grant, L., Yoo, J., Clause, A., Kandler,

K., Noebels, J.L., Glowatzki, E., et al. (2008). Sensorineural deafness and

seizures in mice lacking vesicular glutamate transporter 3. Neuron 57,

263–275.

Smith, R.J., Bale, J.F., Jr., andWhite, K.R. (2005). Sensorineural hearing loss in

children. Lancet 365, 879–890.

Staecker, H., Praetorius, M., Baker, K., and Brough, D.E. (2007). Vestibular hair

cell regeneration and restoration of balance function induced by math1 gene

transfer. Otol. Neurotol. 28, 223–231.

Steven Colburn, H., Shinn-Cunningham, B., Kidd, G., Jr., and Durlach, N.

(2006). The perceptual consequences of binaural hearing. Int. J. Audiol. 45

(Suppl 1 ), S34–S44.

Takamori, S., Malherbe, P., Broger, C., and Jahn, R. (2002). Molecular cloning

and functional characterization of human vesicular glutamate transporter 3.

EMBO Rep. 3, 798–803.

von Ilberg, C.A., Baumann, U., Kiefer, J., Tillein, J., and Adunka, O.F. (2011).

Electric-acoustic stimulation of the auditory system: a review of the first

decade. Audiol. Neurootol. 16 (Suppl 2 ), 1–30.

Neuron 75, 283–293, July 26, 2012 ª2012 Elsevier Inc. 293