Skip to main content
The Journal of Physiology logoLink to The Journal of Physiology
. 2006 Jul 27;576(Pt 1):37–42. doi: 10.1113/jphysiol.2006.114652

Prestin and the cochlear amplifier

Peter Dallos 1, Jing Zheng 1, Mary Ann Cheatham 1
PMCID: PMC1995634  PMID: 16873410

Abstract

In non-mammalian, hair cell-bearing sense organs amplification is associated with mechano-electric transducer channels in the stereovilli (commonly called stereocilia). Because mammals possess differentiated outer hair cells (OHC), they also benefit from a novel electromotile process, powered by the motor protein, prestin. Here we consider new work pertaining to this protein and its potential role as the mammalian cochlear amplifier.


Some form of mechanical amplification is a general property of auditory organs (Manley, 2001; Gopfert et al. 2005). In fact, all hair cell-possessing vertebrates express an ancient universal mechanism, based on mechanical feedback by stereovilli (Martin & Hudspeth, 1999; Martin et al. 2000). A second amplifier arose based on OHC somatic electromotility (Brownell et al. 1985; Kachar et al. 1986; Ashmore, 1987). Although the basic phenomenon of electromotility is well known (Holley, 1996), one of the significant remaining issues pertains to the appropriate partitioning of mammalian amplification between its stereovillar and somatic components.

The nature of prestin

When molecular motors are discussed, attention is usually focused on the linear motors myosin, kinesin and dynein. These cytoskeletal proteins generate motion by using nucleoside triphosphate hydrolysis to produce conformational changes in proteins. Other motors, such as helicases, ribosomal motors, chaperonins, etc., also require ATP hydrolysis for energy. Prestin, however, is different. Because it is a direct voltage-to-force converter, it is probably unique in the animal kingdom. Its closest functional relative is the bacterial flagellum, which is driven by a high-speed rotary motor, dependent on proton current across the bacterial cell membrane (Berg, 1975).

Relying upon the knowledge that only OHCs express the putative gene that codes for the motor protein, we identified prestin (Zheng et al. 2000) using suppression subtractive hybridization PCR (Diatchenko et al. 1996). Prestin's amino acid sequence shows that it belongs to an anion-transporter family, solute carrier protein 26 (SLC26A). Prestin shares the family structure, including a highly conserved central core of hydrophobic amino acids (∼400 a.a.) predicted to form 10–12 transmembrane domains, as well as cytoplasmic N- (∼100 a.a.) and C-termini (∼240 a.a.). Although many members of SLC26A are glycoproteins, glycosylation/deglycosylation does not affect function (Karniski et al. 1998; Matsuda et al. 2004). Although prestin is most closely related to SLC26A6 (Dallos & Fakler, 2002; Mount & Romero, 2004), the human and mouse orthologues of A6 have only 78% amino acid identity. In contrast, prestin is a highly conserved protein with 92.7% of amino acids being identical among four different mammalian species: human, mouse, rat and gerbil (He et al. 2006). Such a high degree of conservation is not common among other SLC26A members. Significant changes in prestin's primary sequence occurred after the split between mammalian and avian lines, suggesting that prestin evolved in order to fit special mammalian needs (Franchini & Elgoyhen, 2006).

Aside from its highly conserved primary sequence, mammalian prestin also has several unique features that differ from other members of the family. First, prestin expresses voltage-dependent charge movement and motility. Second, prestin is abundantly expressed in OHCs (> 107 per cell). Third, prestin exists as stable tetramers (Zheng et al. 2006), which probably form the 11 nm particles observed in the OHC's lateral membrane (Forge, 1991; Kalinec et al. 1992). Although there is little evidence suggesting the existence of oligomeric forms in other members, a possible multimeric form of pendrin was reported, even though the majority of pendrin molecules exist as monomers (Porra et al. 2002). Fourth, although the basic function of SLC26A members is to transport anions (Dallos & Fakler, 2002; Mount & Romero, 2004), this is not prestin's principal role. Recent theoretical work suggests, however, that it is best to model prestin as an electrogenic anion exchanger (Muallem & Ashmore, 2006). Unequivocal experimental verification of anion transport is not yet available.

Early work indicated that prestin is voltage dependent (Santos-Sacchi & Dilger, 1988) and, in analogy with voltage-gated ion channels, it was subsequently assumed that the voltage sensor is a distinct charged moiety displaced across the membrane in response to a change in membrane potential. This displacement presumably initiates a conformational change in the motor component of the protein, producing an alteration of surface area. Charge displacement is manifest as a transient current at the onset and cessation of membrane-potential steps. The total charge displaced at a given potential is the product of the total displaceable charge (Qmax) and the probability (pS) that the displaced charge is in one of its two principal states. This behaviour is expressed as a Boltzmann function. Inasmuch as the charge displaced is voltage dependent, a nonlinear capacitance (NLC) can be associated with it, according to the relationship C = dQ/dV (Ashmore, 1990). Nonlinear capacitance is easier to measure than motility, and it has been shown to provide a surrogate representation (Santos-Sacchi, 1991).

It is reasonable to assume that the voltage-sensing charged group is different for prestin than in other SLC26A proteins, which produce no motility. With this in mind, Oliver et al. (2001) altered each charged, non-conserved amino acid in the putative membrane-interacting region of prestin, individually or in groups. Surprisingly, no combination of mutations eliminated NLC or altered its gain. These results led to the suggestion that the voltage sensor may not be an intrinsic component of the protein, but an extrinsic ion. Using inside-out and outside-out membrane patches, Oliver et al. demonstrated that intracellular Cl functions as the extrinsic voltage sensor. Subsequent investigations (Fakler & Oliver, 2003; Rybalchenko & Santos-Sacchi, 2003; Santos-Sacchi et al. 2006) showed that as intracellular Cl concentration decreases, the amount of charge transferred also decreases and voltage sensitivity shifts in the depolarizing direction. The direction of shift implies that the net charge moved across the membrane is positive. Thus, two alternatives exist to the idea that Cl is the voltage sensor. It is possible that monovalent anions need to attach to a binding site and their combination, with net positivity, is translocated across the membrane. Alternatively, chloride binding could enable an allosteric change, thereby allowing a positive gating charge to be moved (Rybalchenko & Santos-Sacchi, 2003). The two alternatives are demonstrated in the cartoons of Fig. 1.

Figure 1. Two models of prestin gating by voltage, in which the presence of intracellular chloride is an essential factor in both, but the gating mechanisms are different.

Figure 1

The long (or extended) state of the molecule corresponds to hyperpolarization of the cell; the short (or compact) state to depolarization. The no-chloride case is arbitrarily modelled as long. In A, Cl is assumed to associate with a positive binding site and the combination is translocated across the membrane. In B, chloride binding enables a positive gating particle to unlock and be translocated. The cartoons depict incomplete transporters. If full anion transport is demonstrated, a pore region needs to be incorporated.

It was also shown (He & Dallos, 1999, 2000) that voltage-driven motility is accompanied by a voltage-dependent change in axial stiffness. Voltage change can modulate cell stiffness over a range of about 10-fold, with a potential to influence micromechanics. He et al. (2003) also demonstrated that interference with prestin's function, via substitution of pentane sulphonate for chloride, reduces cell stiffness to ∼1/3 of its normal value. Because stiffness changes occur on a cycle-by-cycle basis without significant attenuation as frequency increases, it is conceivable that stiffness change may be the primary voltage-dependent process in OHCs. If so, electromotility would be a simple consequence of changing the stiffness of a preloaded ‘spring’.

Finally, prestin, like other transducers, demonstrates reciprocity (Iwasa, 1993), which can be modelled by assessing the molecule's piezoelectric properties. When viewed in this context, the efficiency of conversion from mechanical force to electrical charge is ∼20 fC nN−1. This number is four orders of magnitude greater than that obtained for the best man-made material (Dong et al. 2002).

Prestin mutations

In order to understand the mechanism underlying electromotility, various mutations, including point mutants, chimeric mutants (hybrid proteins made from portions of prestin and other SLC26A members) and truncation mutants have been investigated (He et al. 2006). Almost all mutants fall into two groups. For group one, NLC is maintained with or without some shift in V1/2, the membrane potential at which half of the molecules are in the contracted state. Group-two mutants lose NLC, in most cases due to their inability to insert into the plasma membrane (PM). Like other members of the SLC26A family, prestin contains a sulphate transporter and antisigma-factor antagonist (STAS) domain, the latter located in the C-terminus. Mutations occurring in this region result in improper targeting and loss of partial or complete function (Taylor et al. 2002; Karniski, 2004). Prestin is no exception (Zheng et al. 2005).

Mutations in other regions of prestin can also cause improper PM targeting. In a heterologous system, prestin mutants are often misfolded or aggregated, resulting in their retention and accumulation in the endoplasmic reticulum and other cytoplasmic membranes. In fact, deletion of more than 21 a.a. at the N terminus, or more than 32 a.a. at the C terminus, results in loss of function, probably due to improper targeting (Navaratnam et al. 2005; Zheng et al. 2005). Other family members demonstrate similar problems, such as Pendred syndrome (Rotman-Pikielny et al. 2002; Taylor et al. 2002). In fact, there is a strong negative correlation between the severity of the human disease phenotype and the level of SLC26A protein present in the PM (Scott et al. 2000; Karniski, 2004).

Prestin as amplifier

It is known that the low-pass cutoff frequency of the OHC membrane filter is at most ∼1 kHz (Housley & Ashmore, 1992; Preyer et al. 1996). Thus, the cell's receptor potential, assumed to control motility, is attenuated above the cutoff frequency and is progressively less effective. In contrast, the motor itself is fast, i.e. OHC motility can be induced at greater than 70 kHz (Frank et al. 1999). In order to envision how motility occurs in vivo, ingenious schemes have been proposed to overcome the filter problem. Among these are various models utilizing prestin's piezoelectric properties to compensate for the voltage attenuation (Mountain & Hubbard, 1994; Dong et al. 2002; Ospeck et al. 2003; Spector et al. 2003), interaction of prestin with localized chloride conductances (Rybalchenko & Santos-Sacchi, 2003) and reliance on extracellular voltage gradients (Dallos & Evans, 1995). This latter conjecture has been tested experimentally (Fridberger et al. 2004) and found to merit consideration. Finally, the recognition that cochlear amplification can be construed as a negative feedback process (Mountain & Hubbard, 1994) was used to demonstrate that while individual OHCs may have low-pass filter properties, the resulting system bandwidth is much wider when their collective gain is high (Lu et al. 2006). Although it is not known which of the above possibilities, or their combinations, might be employed in the living cochlea, there clearly are several plausible means whereby the bandwidth limitation can be overcome.

Prestin knockout mice

Measurements in prestin knockout (KO) mice demonstrate that prestin is required for normal auditory function. In the absence of prestin, OHCs do not exhibit electromotility (Liberman et al. 2002). In vivo, brain-stem evoked-potential thresholds increase by ∼50 dB (Liberman et al. 2004) as do compound action potential thresholds (Cheatham et al. 2004). Frequency selectivity is also absent (Cheatham et al. 2004). In order to associate these physiological deficits with a change in OHC motor function, it must be established that forward transduction is normal in KO mice. Homozygotes have wildtype-like nonlinear responses including harmonic and intermodulation distortion (Liberman et al. 2004; Cheatham et al. 2004, 2006), CM pseudotransducer functions (peak ± ac (alternating current) response versus peak ± sound pressure), both summating potential polarities, as well as normal uptake of the dye AM1-43 via transducer channels (Cheatham et al. 2004). Data from OHCs isolated from KO mice also demonstrate large asymmetrical transducer currents similar to those in wildtype controls (Jia et al. 2006). Thus mechano-electrical transduction appears to be normal in mice lacking prestin and therefore changes in sensitivity and frequency selectivity cannot be attributed to changes in forward transduction. Because prestin KO mice exhibit a progressive apoptosis of both inner and outer hair cells (Liberman et al. 2002; Wu et al. 2004), older adult mice are probably better models for deaf adult humans with prestin mutations (Liu et al. 2003). Data also suggest that heterozygous prestin mice have normal sensitivity and frequency selectivity, as well as near normal levels of prestin protein and OHC somatic electromotility (Liberman et al. 2004; Cheatham et al. 2005). Because prestin mRNA is ∼50% of wildtype (Liberman et al. 2002), prestin protein expression appears to be autoregulated (Cheatham et al. 2005).

Somatic and stereovillar amplification

The stereovillar motor can do work against an external load (Benser et al. 1996), may have the necessary speed (Kennedy et al. 2005) and can produce some nonlinear amplification (Chan & Hudspeth, 2005). However, there is no evidence that it dominates mammalian amplification. Because non-mammalian sense organs lack OHCs and prestin, it is assumed that the stereovillar amplifier is their principal means of boosting hearing sensitivity, while the newer somatic motility evolved to fulfil some specific mammalian requirement, possibly associated with their extended frequency range of hearing. If stereovillar amplification were the primary mechanism throughout this range, the evolutionary advantage of adding somatic motility becomes obscure. Hence, we assume that both mechanisms are present in mammals. Although de novo somatic motility probably dominates, spontaneous otoacoustic emissions may be more intimately tied to stereovillar mechanisms. In this context, we note reports showing spontaneous oscillations of stereovillar bundles (Crawford & Fettiplace, 1985; Rusch & Thurm, 1990; Martin & Hudspeth, 1999; Camalet et al. 2000). In contrast, no spontaneous oscillation has ever been seen in OHC length or stiffness.

Conclusions

Remaining problems include an experimental examination of the molecule's putative transport function, its means of voltage sensing, its companion proteins, the nature of its conformational change and its possible interactions with surrounding lipid (Raphael et al. 2000; Zhang et al. 2001). Many of these questions may be answered when a structural description becomes available. The most important outstanding issue, however, is the apportioning of mammalian amplification between stereovillar and somatic motility (Chan & Hudspeth, 2005; Jia et al. 2006; Kennedy et al. 2006).

Acknowledgments

Work supported by NIDCD grant DC00089.

References

  1. Ashmore JF. A fast motile response in guinea-pig outer hair cells: the cellular basis of the cochlear amplifier. J Physiol. 1987;388:323–347. doi: 10.1113/jphysiol.1987.sp016617. [DOI] [PMC free article] [PubMed] [Google Scholar]
  2. Ashmore JF. Forward and reverse transduction in the mammalian cochlea. Neurosci Res. 1990;12(Suppl):S39–S50. doi: 10.1016/0921-8696(90)90007-p. [DOI] [PubMed] [Google Scholar]
  3. Benser ME, Marquis RE, Hudspeth AJ. Rapid, active hair bundle movements in hair cells from the bullfrog's sacculus. J Neurosci. 1996;16:5629–5643. doi: 10.1523/JNEUROSCI.16-18-05629.1996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  4. Berg HC. Chemotaxis in bacteria. Ann Rev Biophys Bioeng. 1975;4:119–136. doi: 10.1146/annurev.bb.04.060175.001003. [DOI] [PubMed] [Google Scholar]
  5. Brownell WE, Bader CR, Bertrand D, De Ribaupierre Y. Evoked mechanical responses of isolated cochlear outer hair cells. Science. 1985;227:194–196. doi: 10.1126/science.3966153. [DOI] [PubMed] [Google Scholar]
  6. Camalet S, Duke T, Julicher F, Prost J. Auditory sensitivity provided by self-tuned critical oscillations of hair cells. Proc Natl Acad Sci U S A. 2000;97:3183–3188. doi: 10.1073/pnas.97.7.3183. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Chan DK, Hudspeth AJ. Ca2+ current-driven nonlinear amplification by the mammalian cochlea in vitro. Nat Neurosci. 2005;8:149–155. doi: 10.1038/nn1385. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Cheatham MA, Huynh KH, Dallos P. Nonlinear responses in prestin knockout mice: Implications for cochlear function. In: Nuttall AL, editor. Auditory Mechanisms: Processes and Models. Singapore: World Scientific; 2006. pp. 311–318. Portland OR. [Google Scholar]
  9. Cheatham MA, Huynh KH, Gao J, Zuo J, Dallos P. Cochlear function in Prestin knockout mice. J Physiol. 2004;560:821–830. doi: 10.1113/jphysiol.2004.069559. [DOI] [PMC free article] [PubMed] [Google Scholar]
  10. Cheatham MA, Zheng J, Huynh KH, Du GG, Gao J, Zuo J, Navarrete E, Dallos P. Cochlear function in mice with only one copy of the prestin gene. J Physiol. 2005;569:229–241. doi: 10.1113/jphysiol.2005.093518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  11. Crawford AC, Fettiplace R. The mechanical properties of ciliary bundles of turtle cochlear hair cells. J Physiol. 1985;364:359–379. doi: 10.1113/jphysiol.1985.sp015750. [DOI] [PMC free article] [PubMed] [Google Scholar]
  12. Dallos P, Evans BN. High-frequency motility of outer hair cells and the cochlear amplifier. Science. 1995;267:2006–2009. doi: 10.1126/science.7701325. [DOI] [PubMed] [Google Scholar]
  13. Dallos P, Fakler B. Prestin, a new type of motor protein. Nat Rev Mol Cell Biol. 2002;3:104–111. doi: 10.1038/nrm730. [DOI] [PubMed] [Google Scholar]
  14. Diatchenko L, Lau YF, Campbell AP, Chenchik A, Moqadam F, Huang B, Lukyanov S, Lukyanov K, Gurskaya N, Sverdlov ED, Siebert PD. Suppression subtractive hybridization: a method for generating differentially regulated or tissue-specific cDNA probes and libraries. Proc Natl Acad Sci U S A. 1996;93:6025–6030. doi: 10.1073/pnas.93.12.6025. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Dong XX, Ospeck M, Iwasa KH. Piezoelectric reciprocal relationship of the membrane motor in the cochlear outer hair cell. Biophys J. 2002;82:1254–1259. doi: 10.1016/S0006-3495(02)75481-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
  16. Fakler B, Oliver D. Functional properties of prestin – How the motor molecule works. In: Gummer AW, editor. Biophysics of the Cochlea from Molecules to Models. London: World Scientific; 2003. pp. 110–114. [Google Scholar]
  17. Forge A. Structural features of the lateral walls in mammalian cochlear outer hair cells. Cell Tissue Res. 1991;265:473–483. doi: 10.1007/BF00340870. [DOI] [PubMed] [Google Scholar]
  18. Franchini LF, Elgoyhen AB. Adaptive evolution in mammalian proteins involved in cochlear outer hair cell electromotility. Molec Phylogen Evol. 2006 doi: 10.1016/j.ympev.2006.05.042. in press. [DOI] [PubMed] [Google Scholar]
  19. Frank G, Hemmert W, Gummer AW. Limiting dynamics of high-frequency electromechanical transduction of outer hair cells. Proc Natl Acad Sci U S A. 1999;96:4420–4425. doi: 10.1073/pnas.96.8.4420. [DOI] [PMC free article] [PubMed] [Google Scholar]
  20. Fridberger A, De Monvel JB, Zheng J, Hu N, Zou Y, Ren T, Nuttall A. Organ of corti potentials and the motion of the basilar membrane. J Neurosci. 2004;24:10057–10063. doi: 10.1523/JNEUROSCI.2711-04.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Gopfert MC, Humphris AD, Albert JT, Robert D, Hendrich O. Power gain exhibited by motile mechanosensory neurons in Drosophila ears. Proc Natl Acad Sci U S A. 2005;102:325–330. doi: 10.1073/pnas.0405741102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  22. He DZ, Dallos P. Somatic stiffness of cochlear outer hair cells is voltage-dependent. Proc Natl Acad Sci U S A. 1999;96:8223–8228. doi: 10.1073/pnas.96.14.8223. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. He DZ, Dallos P. Properties of voltage-dependent somatic stiffness of cochlear outer hair cells. J Assoc Res Otolaryngol. 2000;1:64–81. doi: 10.1007/s101620010006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  24. He DZ, Jia S, Dallos P. Prestin and the dynamic stiffness of cochlear outer hair cells. J Neurosci. 2003;23:9089–9096. doi: 10.1523/JNEUROSCI.23-27-09089.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  25. He DZ, Zheng J, Kalinec F, Kakehata S, Santos-Sacchi J. Tuning in to the amazing outer hair cell: Membrane wizardry with a twist and shout. J Membr Biol. 2006;209:119–134. doi: 10.1007/s00232-005-0833-9. [DOI] [PubMed] [Google Scholar]
  26. Holley MC. In: The Cochlea: The Springer Handbook of Auditory Research. Dallos P, Popper AN, Fay RF, editors. Vol. 8. New York: Springer; 1996. pp. 386–434. [Google Scholar]
  27. Housley GD, Ashmore JF. Ionic currents of outer hair cells isolated from the guinea-pig cochlea. J Physiol. 1992;448:73–98. doi: 10.1113/jphysiol.1992.sp019030. [DOI] [PMC free article] [PubMed] [Google Scholar]
  28. Iwasa KH. Effect of stress on the membrane capacitance of the auditory outer hair cell. Biophys J. 1993;65:492–498. doi: 10.1016/S0006-3495(93)81053-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
  29. Jia S, Zuo J, Dallos P, He DZZ. The cochlear amplifier: Is it hair bundle motion of outer hair cells? In: Nuttall AF, editor. Auditory Mechanisms: Processes and Models. Singapore: World Scientific; 2006. pp. 201–208. Portland OR. [Google Scholar]
  30. Kachar B, Brownell WE, Altschuler R, Fex J. Electrokinetic shape changes of cochlear outer hair cells. Nature. 1986;322:365–368. doi: 10.1038/322365a0. [DOI] [PubMed] [Google Scholar]
  31. Kalinec F, Holley MC, Iwasa KH, Lim DJ, Kachar B. A membrane-based force generation mechanism in auditory sensory cells. Proc Natl Acad Sci U S A. 1992;89:8671–8675. doi: 10.1073/pnas.89.18.8671. [DOI] [PMC free article] [PubMed] [Google Scholar]
  32. Karniski LP. Functional expression and cellular distribution of diastrophic dysplasia sulfate transporter (DTDST) gene mutations in HEK cells. Hum Mol Genet. 2004;13:2165–2171. doi: 10.1093/hmg/ddh242. [DOI] [PubMed] [Google Scholar]
  33. Karniski LP, Lotscher M, Fucentese M, Hilfiker H, Biber J, Murer H. Immunolocalization of sat-1 sulfate/oxalate/bicarbonate anion exchanger in the rat kidney. Am J Physiol. 1998;275:F79–F87. doi: 10.1152/ajprenal.1998.275.1.F79. [DOI] [PubMed] [Google Scholar]
  34. Kennedy HJ, Crawford AC, Fettiplace R. Force generation by mammalian hair bundles supports a role in cochlear amplification. Nature. 2005;433:880–883. doi: 10.1038/nature03367. [DOI] [PubMed] [Google Scholar]
  35. Kennedy HJ, Evans MG, Crawford AC, Fettiplace R. Depolarization of cochlear outer hair cells evokes active hair bundle motion by two mechanisms. J Neurosci. 2006;26:2757–2766. doi: 10.1523/JNEUROSCI.3808-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  36. Liberman MC, Gao J, He DZ, Wu X, Jia S, Zuo J. Prestin is required for electromotility of the outer hair cell and for the cochlear amplifier. Nature. 2002;419:300–304. doi: 10.1038/nature01059. [DOI] [PubMed] [Google Scholar]
  37. Liberman MC, Zuo J, Guinan JJ., Jr Otoacoustic emissions without somatic motility: can stereocilia mechanics drive the mammalian cochlea. J Acoust Soc Am. 2004;116:1649–1655. doi: 10.1121/1.1775275. [DOI] [PMC free article] [PubMed] [Google Scholar]
  38. Liu XZ, Ouyang XM, Xia XJ, Zheng J, Pandya A, Li F, Du LL, Welch KO, Petit C, Smith RJ, Webb BT, Yan D, Arnos KS, Corey D, Dallos P, Nance WE, Chen ZY. Prestin, a cochlear motor protein, is defective in non-syndromic hearing loss. Hum Mol Genet. 2003;12:1155–1162. doi: 10.1093/hmg/ddg127. [DOI] [PubMed] [Google Scholar]
  39. Lu TK, Zhak S, Dallos P, Sarpeshkar R. Fast cochlear amplification with slow outer hair cells. Hear Res. 2006;214:45–67. doi: 10.1016/j.heares.2006.01.018. [DOI] [PubMed] [Google Scholar]
  40. Manley GA. Evidence for an active process and a cochlear amplifier in nonmammals. J Neurophysiol. 2001;86:541–549. doi: 10.1152/jn.2001.86.2.541. [DOI] [PubMed] [Google Scholar]
  41. Martin P, Hudspeth AJ. Active hair-bundle movements can amplify a hair cell's response to oscillatory mechanical stimuli. Proc Natl Acad Sci U S A. 1999;96:14306–14311. doi: 10.1073/pnas.96.25.14306. [DOI] [PMC free article] [PubMed] [Google Scholar]
  42. Martin P, Mehta AD, Hudspeth AJ. Negative hair-bundle stiffness betrays a mechanism for mechanical amplification by the hair cell. Proc Natl Acad Sci U S A. 2000;97:12026–12031. doi: 10.1073/pnas.210389497. [DOI] [PMC free article] [PubMed] [Google Scholar]
  43. Matsuda K, Zheng JG, Du G, Klocker N, Madison LD, Dallos P. N-linked glycosylation sites of the motor protein prestin: effects on membrane targeting and electrophysiological function. J Neurochem. 2004;89:928–938. doi: 10.1111/j.1471-4159.2004.02377.x. [DOI] [PubMed] [Google Scholar]
  44. Mount DB, Romero MF. The SLC26 gene family of multifunctional anion exchangers. Pflugers Arch. 2004;447:710–721. doi: 10.1007/s00424-003-1090-3. [DOI] [PubMed] [Google Scholar]
  45. Mountain DC, Hubbard AE. A piezoelectric model of outer hair cell function. J Acoust Soc Am. 1994;95:350–354. doi: 10.1121/1.408273. [DOI] [PubMed] [Google Scholar]
  46. Muallem D, Ashmore J. An anion antiporter model of prestin, the outer hair cell motor protein. Biophys J. 2006;90:4035–4045. doi: 10.1529/biophysj.105.073254. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Navaratnam D, Bai JP, Samaranayake H, Santos-Sacchi J. N-terminal-mediated homomultimerization of prestin, the outer hair cell motor protein. Biophys J. 2005;89:3345–3352. doi: 10.1529/biophysj.105.068759. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Oliver D, He DZ, Klocker N, Ludwig J, Schulte U, Waldegger S, Ruppersberg JP, Dallos P, Fakler B. Intracellular anions as the voltage sensor of prestin, the outer hair cell motor protein. Science. 2001;292:2340–2343. doi: 10.1126/science.1060939. [DOI] [PubMed] [Google Scholar]
  49. Ospeck M, Dong XX, Iwasa KH. Limiting frequency of the cochlear amplifier based on electromotility of outer hair cells. Biophys J. 2003;84:739–749. doi: 10.1016/S0006-3495(03)74893-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
  50. Porra V, Bernier-Valentin F, Trouttet-Masson S, Berger-Dutrieux N, Peix JL, Perrin A, Selmi-Ruby S, Rousset B. Characterization and semiquantitative analyses of pendrin expressed in normal and tumoral human thyroid tissues. J Clin Endocrinol Metab. 2002;87:1700–1707. doi: 10.1210/jcem.87.4.8372. [DOI] [PubMed] [Google Scholar]
  51. Preyer S, Renz S, Hemmert W, Zenner H-P, Gummer AW. Receptor potential of outer hair cells isolated from base to apex of the adult guinea-pig cochlea: Implications for cochlear tuning mechanisms. Auditory Neuroscience. 1996;2:145–157. [Google Scholar]
  52. Raphael RM, Popel AS, Brownell WE. A membrane bending model of outer hair cell electromotility. Biophys J. 2000;78:2844–2862. doi: 10.1016/S0006-3495(00)76827-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
  53. Rotman-Pikielny P, Hirschberg K, Maruvada P, Suzuki K, Royaux IE, Green ED, Kohn LD, Lippincott-Schwartz J, Yen PM. Retention of pendrin in the endoplasmic reticulum is a major mechanism for Pendred syndrome. Hum Mol Genet. 2002;11:2625–2633. doi: 10.1093/hmg/11.21.2625. [DOI] [PubMed] [Google Scholar]
  54. Rusch A, Thurm U. Spontaneous and electrically induced movements of ampullary kinocilia and stereovilli. Hear Res. 1990;48:247–263. doi: 10.1016/0378-5955(90)90065-w. [DOI] [PubMed] [Google Scholar]
  55. Rybalchenko V, Santos-Sacchi J. Cl− flux through a non-selective, stretch-sensitive conductance influences the outer hair cell motor of the guinea-pig. J Physiol. 2003;547:873–891. doi: 10.1113/jphysiol.2002.036434. [DOI] [PMC free article] [PubMed] [Google Scholar]
  56. Santos-Sacchi J. Reversible inhibition of voltage-dependent outer hair cell motility and capacitance. J Neurosci. 1991;11:3096–3110. doi: 10.1523/JNEUROSCI.11-10-03096.1991. [DOI] [PMC free article] [PubMed] [Google Scholar]
  57. Santos-Sacchi J, Dilger JP. Whole cell currents and mechanical responses of isolated outer hair cells. Hear Res. 1988;35:143–150. doi: 10.1016/0378-5955(88)90113-x. [DOI] [PubMed] [Google Scholar]
  58. Santos-Sacchi J, Song L, Zheng J, Nuttall AL. Control of mammalian cochlear amplification by chloride anions. J Neurosci. 2006;26:3992–3998. doi: 10.1523/JNEUROSCI.4548-05.2006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  59. Scott DA, Wang R, Kreman TM, Andrews M, McDonald JM, Bishop JR, Smith RJ, Karniski LP, Sheffield VC. Functional differences of the PDS gene product are associated with phenotypic variation in patients with Pendred syndrome and non- syndromic hearing loss (DFNB4) Hum Mol Genet. 2000;9:1709–1715. doi: 10.1093/hmg/9.11.1709. [DOI] [PubMed] [Google Scholar]
  60. Spector AA, Brownell WE, Popel AS. Effect of outer hair cell piezoelectricity on high-frequency receptor potentials. J Acoust Soc Am. 2003;113:453–461. doi: 10.1121/1.1526493. [DOI] [PubMed] [Google Scholar]
  61. Taylor JP, Metcalfe RA, Watson PF, Weetman AP, Trembath RC. Mutations of the PDS gene, encoding pendrin, are associated with protein mislocalization and loss of iodide efflux: implications for thyroid dysfunction in Pendred syndrome. J Clin Endocrinol Metab. 2002;87:1778–1784. doi: 10.1210/jcem.87.4.8435. [DOI] [PubMed] [Google Scholar]
  62. Wu X, Gao J, Guo Y, Zuo J. Hearing threshold elevation precedes hair-cell loss in prestin knockout mice. Brain Res Mol Brain Res. 2004;126:30–37. doi: 10.1016/j.molbrainres.2004.03.020. [DOI] [PubMed] [Google Scholar]
  63. Zhang PC, Keleshian AM, Sachs F. Voltage-induced membrane movement. Nature. 2001;413:428–432. doi: 10.1038/35096578. [DOI] [PubMed] [Google Scholar]
  64. Zheng J, Du GG, Anderson CT, Keller JP, Orem A, Dallos P, Cheatham M. Analysis of the oligomeric structure of the motor protein prestin. J Biol Chem. 2006;281:19916–19924. doi: 10.1074/jbc.M513854200. [DOI] [PubMed] [Google Scholar]
  65. Zheng J, Gu GG, Matsuda K, Orem A, Aguinaga S, Deak L, Navarrete E, Madison L, Dallos P. The carboxy terminus of prestin influences both motor function and plasma membrane targeting. J Cell Sci. 2005;118:2987–2996. doi: 10.1242/jcs.02431. [DOI] [PubMed] [Google Scholar]
  66. Zheng J, Shen W, He DZ, Long KB, Madison LD, Dallos P. Prestin is the motor protein of cochlear outer hair cells. Nature. 2000;405:149–155. doi: 10.1038/35012009. [DOI] [PubMed] [Google Scholar]

Articles from The Journal of Physiology are provided here courtesy of The Physiological Society

RESOURCES