Hostname: page-component-7c8c6479df-24hb2 Total loading time: 0 Render date: 2024-03-28T19:42:10.876Z Has data issue: false hasContentIssue false

Energy flow in the cochlea

Published online by Cambridge University Press:  20 April 2006

James Lighthill
Affiliation:
University College London

Abstract

With moderate acoustic stimuli, measurements of basilar-membrane vibration (especially, those using a Mössbauer source attached to the membrane) demonstrate:

  1. a high degree of asymmetry, in that the response to a pure tone falls extremely sharply above the characteristic frequency, although much more gradually below it;

  2. a substantial phase-lag in that response, and one which increases monotonically up to the characteristic frequency;

  3. a response to a ‘click’ in the form of a delayed ‘ringing’ oscillation at the characteristic frequency, which persists for around 20 cycles.

This paper uses energy-flow considerations to identify which features in a mathematical model of cochlear mechanics are necessary if it is to reproduce these experimental findings.

The response (iii) demands a travelling-wave model which incorporates an only lightly damped resonance. Admittedly, waveguide systems including resonance are described in classical applied physics. However, a classical waveguide resonance reflects a travelling wave, thus converting it into a standing wave devoid of the substantial phase-lag (ii); and produces a low-frequency cutoff instead of the high-frequency cutoff (i).

By contrast, another general type of travelling-wave system with resonance has become known more recently; initially, in a quite different context (physics of the atmosphere). This is described as critical-layer resonance, or else (because the resonance absorbs energy) critical-layer absorption. It yields a high-frequency cutoff; but, above all, it is characterized by the properties of the energy flow velocity. This falls to zero very steeply as the point of resonance is approached; so that wave energy flow is retarded drastically, giving any light damping which is present an unlimited time in which to dissipate that energy.

Existing mathematical models of cochlear mechanics, whether using one-, two- or three-dimensional representations of cochlear geometry, are analysed from this standpoint. All are found to have been successful (if only light damping is incorporated, as (iii) requires) when and only when they incorporate critical-layer absorption. This resolves the paradox of why certain grossly unrealistic one-dimensional models can give a good prediction of cochlear response; it is because they incorporate the one essential feature of critical-layer absorption.

At any point in a physical system, the high-frequency limit of energy flow velocity is the slope of the graph of frequency against wavenumberIn any travelling wave, the wavenumber is the rate of change of phase with distance; for example, it is 2π/λ in a sine wave of length λ. at that point. In the cochlea, this is a good approximation at frequencies above about 1 kHz; and, even at much lower frequencies, remains good for wavenumbers above about 0·2 mm−1 (which excludes only a relatively unimportant region near the base).

Frequency of vibration at any point can vary with wavenumber either because stiffness or inertia varies with wavenumber. However, we find that models incorporating a wavenumber-dependent membrane stiffness must be abandoned because they fail to give critical-layer absorption; this is why their predictions (when realistically light damping is used) have been unsuccessful. Similarly, models neglecting the inertia of the cochlear partition must be rejected.

One-dimensional modelling becomes physically unrealistic for wavenumbers above about 0.7 mni-1, and the error increases with wavenumber. The main trouble is that a one-dimensional theory makes the effective inertia ‘flatten out’ to its limiting value (inertia of the cochlear partition alone) too rapidly as wavenumber increases. Fortunately, a two-dimensional, or even a three-dimensional, model can readily be used to calculate a more realistic, and significantly more gradual, ‘flattening out’ of this inertia. All of the models give a fair representation of the experimental data, because they all predict critical-layer absorption. However, the more realistic two- or three-dimensional models must be preferred. These retard the wave energy flow still more, thus facilitating its absorption by even a very modest level of damping. The paper indicates many other features of these models.

The analysis described above is preceded by a discussion of waves generated a t the oval window. They necessarily include:

  1. the already-mentioned travelling wave, or ‘slow wave’, in which the speed of energy flow falls from around 100 m s−l at the base to zero at the position of resonance;

  2. a pure sound wave, or ‘fast wave’, travelling a t 1400 m s-1, with reflection at the apex which makes i t into a standing wave. Half of the rate of working of the stapes footplate against the oval window is communicated as an energy flow a t this much higher speed down the scala vestibuli, across the cochlear partition and back up the scala tympani to the round window, whence it becomes part of the slow general apical progress of the travelling wave; a progress which, as described above, comes to a halt altogether just in front of the position of resonance.

Mathematical detail is avoided in the discussion of cochlear energy flow in the main part of the paper (§ 1–10), but a variety of relevant mathematical analysis is given in appendices A–E. These include, also, new comments about the functions of the tunnel of Corti (appendix A) and the helicotrema (appendix C).

Type
Research Article
Copyright
© 1981 Cambridge University Press

Access options

Get access to the full version of this content by using one of the access options below. (Log in options will check for institutional or personal access. Content may require purchase if you do not have access.)

References

Allen, J. B. 1977 Two-dimensional cochlear fluid model: new results. J. Acoust. Soc. Am. 61, 110119.Google Scholar
Békésy, G. von 1941 Über die Elastizität der Schneckentrennwand des Ohres. Akust. Zeits. 6, 265278.Google Scholar
Békésy, G. von 1947 The variation of phase along the basilar membrane with sinusoidal vibrations. J. Acoust. Soc. Am. 19, 452460.Google Scholar
Békésy, G. von 1960 Experiments in Hearing. McGraw-Hill.
Berkley, D. A. & Lesser, M. B. 1973 Comparison of single- and double-chamber models of the cochlea. J. Acoust. Soc. Am. 53, 10371038.Google Scholar
Boer, E. de 1979 Short-wave world revisited: resonance in a two-dimensional cochlear model. Hearing Res. 1, 253258.Google Scholar
Bogert, B. P. 1950 Determination of the effects of dissipation in the cochlear partition by means of a network representing the basilar membrane. J. Acoust. Soc. Am. 23, 151154.Google Scholar
Booker, J. R. & Bretherton, F. P. 1967 The critical layer for internal gravity waves in a shear flow. J. Fluid Mech. 27, 513539.Google Scholar
Cole, J. D. & Chadwick, R. S. 1977 An approach to the mechanics of the cochlea. Z. angew. Math. Phys. 28, 785804.Google Scholar
Evans, E. F. & Wilson, J. P. 1973 The frequency selectivity of the cochlea. In Basic Mechanisms in Hearing (ed. A. R. Möller), pp. 519554. Academic.
Fettiplace, R. & Crawford, A. C. 1978 The coding of sound pressure and frequency in cochlear hair cells of the terrapin. Proc. Roy. Soc. B 203, 209218.Google Scholar
Fletcher, H. 1951 On the dynamics of the cochlea. J. Acoust. Soc. Am. 23, 637645.Google Scholar
Geisler, C. D. 1976 Mathematical models of the mechanics of the inner ear. In Handbook of Sensory Physiology, vol. v/3. Auditory System: Clinical and Special Topics (ed. W. D. Keidel & W. D. Neff), pp. 391415. Springer.
Greenwood, D. D. 1961 Critical bandwidth and the frequency co-ordinates of the basilar membrane. J. Acoust. Soc. Am. 33, 13441357.Google Scholar
Gundersen, T., Skarstein, O. & Sikkeland, T. 1978 A study of the vibration of the basilar membrane in human temporal bone preparations by the use of the Mössbauer effect. Acta Otolaryngol. 86, 225232.Google Scholar
Huxley, A. F. 1969 Is resonance possible in the cochlea after all? Nature 221, 935940.Google Scholar
Johnstone, B. M., Taylor, K. J. & Boyle, A. J. 1970 Mechanics of the guinea pig cochlea. J. Acoust. Soc. Am. 47, 504509.Google Scholar
Kiang, N. Y. S. & Moxon, E. C. 1974 Tails of tuning curves of auditory-nerve fibres. J. Acoust. Soc. Am. 55, 620630.Google Scholar
Kiang, N. Y. S., Watanabe, T., Thomas, E. C. & Clark, L. F. 1965 Discharge Patterns of Single Fibres in the Cat's Auditory Nerve. Massachusetts Institute of Technology Press.
Kohllöffel, W. E. 1972 A study of basilar-membrane vibrations. Acustica 27, 4989.Google Scholar
Kohllöffel, W. E. 1973 Observations of the mechanical disturbances along the basilar membrane with laser illumination. In Basic Mechanisms in Hearing (ed. A. R. Möller), pp. 95113. Academic.
Laming, D. R. J. 1979 Unpublished work, to appear shortly as Chapter 8 of Weber's Law. Academic.
Lesser, M. B. & Berkley, D. A. 1972 Fluid mechanics of the cochlea, part 1. J. Fluid Mech. 51, 497512.Google Scholar
Lighthill, J. 1978 Waves in Fluids. Cambridge University Press.
Peterson, L. C. & Bogert, B. P. 1950 A dynamical theory of the cochlea. J. Acoust. Soc. Am. 22, 369381.Google Scholar
Ranke, O. F. 1950 Hydrodynamik der Schneckenflüssigkeit. Zeits. f. Biol. 103, 409434.Google Scholar
Rhode, W. S. 1971 Observation of the vibration of the basilar membrane in squirrel monkeys using the Mössbauer technique. J. Acoust. Soc. Am. 49, 12181231.Google Scholar
Rhode, W. S. 1973 An investigation of post-mortem cochlear mechanics using the Mössbauer effect. In Basic Mechanisms in Hearing (ed. A. R. Möller), pp. 4963. Academic.
Robles, L., Rhode, W. S. & Geisler, C. D. 1976 Transient response of the basilar membrane measured in squirrel monkeys using the Mössbauer effect. J. Acoust. Soc. Am. 59, 926939.Google Scholar
Schroeder, M. R. 1973 An integrable model for the basilar membrane. J. Acoust. Soc. Am. 53, 429434.Google Scholar
Siebert, W. M. 1974 Ranke revisited—a simple short-wave cochlear model. J. Acoust. Soc. Am. 56, 594600.Google Scholar
Steele, C. R. 1974 Behaviour of the basilar membrane with pure-tone excitation. J. Acoust. Soc. Am. 55, 148162.Google Scholar
Steele, C. R. 1976 Cochlear mechanics. In Handbook of Sensory Physiology, vol. v/3, Auditory System: Clinical and Special Topics (ed. W. D. Keidel & W. D. Neff), pp. 443478. Springer.
Steele, C. R. & Taber, L. A. 1979a Comparison of ‘WKB’ and finite difference calculations for a two-dimensional cochlear model. J. Acoust. Soc. Am. 65, 10011006.Google Scholar
Steele, C. R. & Taber, L. A. 1979b Comparison of WKB calculations and experimental results for three-dimensional cochlear models. J. Acoust. Soc. Am. 65, 10071018.Google Scholar
Viergever, M. 1980 Mechanics of the Inner Ear. Delft University Press.
Voldřich, L. 1978 Mechanical properties of basilar membrane. Acta Otolaryngol. 86, 331335.Google Scholar
Wever, E. G. & Lawrence, M. 1954 Physiological Acoustics. Princeton University Press.
Wilson, J. P. & Johnstone, J. R. 1975 Basilar membrane and middle-ear vibration in guinea pig measured by capacitance probe. J. Acoust. Soc. Am. 57, 705723.Google Scholar
Zweig, G. 1976 Basilar membrane motion. Cold Spring Harbor Symposia on Quantit. Biol. 40, 619633.Google Scholar
Zweig, G., Lipes, R. & Pierce, J. R. 1976 The cochlear compromise. J. Acoust. Soc. Am. 59, 975982.Google Scholar
Zwislocki, J. 1948 Theorie der Schneckenmechanik. Acta Otolaryngol. Suppl. 72.Google Scholar
Zwislocki, J. 1965 Analysis of some auditory characteristics. In Handbook of Mathematical Psychology (ed. R. D. Luce, R. R. Bush & E. Galanter), vol. III, pp. 197. Wiley.