Possible mechanisms of cochlear two-tone suppression represented by vector subtraction within a model

Similar documents
Mechanical Properties of the Cochlea. Reading: Yost Ch. 7

Effect of the Attachment of the Tectorial Membrane on Cochlear Micromechanics and Two-Tone Suppression

Comparing in vitro, in situ, and in vivo experimental data in a three-dimensional model of mammalian cochlear mechanics

A truly remarkable aspect of human hearing is the vast

An active cochlear model showing sharp tuning and high sensitivity

HST 721 Lecture 4: Mechanics, electromotility and the cochlear amplifier

Half-Octave Shift in Mammalian Hearing Is an Epiphenomenon of the Cochlear Amplifier

THE INTERPLAY BETWEEN ACTIVE HAIR BUNDLE MECHANICS AND ELECTROMOTILITY IN THE COCHLEA

Chapter 3: Anatomy and physiology of the sensory auditory mechanism

Acoustics Research Institute

A cochlear AGC model, proposing a new type of cochlear non-linearity

Auditory nerve model for predicting performance limits of normal and impaired listeners

P T P V U V I. INTRODUCTION. II. MODEL FORMULATION A. Cochlear fluid dynamics.

Two-Tone Suppression of Simultaneous Electrical and Mechanical Responses in the Cochlea

HST 721 Efferent Control Lecture October 2004

Reticular lamina and basilar membrane vibrations in living mouse cochleae

Representation of sound in the auditory nerve

Processing of sounds in the inner ear

Auditory System. Barb Rohrer (SEI )

Improving the diagnostic power of otoacoustic emissions. Arturo Moleti Physics Department University of Roma Tor Vergata

FIFTH INTERNATIONAL CONGRESS ON SOUND AND VIBRATION DECEMBER 15-18, 1997 ADELAIDE, SOUTH AUSTRALIA. Invited Paper LINEAR RESPONSE OF THE COCHLEA

sensitivity (db) frequency (khz) I: MONGOLIAN GERBIL FREQUENCY HEARING CHARACTERISTICS

Harmonic distortion on the basilar membrane in the basal turn of the guinea-pig cochlea

I. INTRODUCTION. J. Acoust. Soc. Am. 111 (1), Pt. 1, Jan /2002/111(1)/271/14/$ Acoustical Society of America

A computer model of medial efferent suppression in the mammalian auditory system

Study the Effect of the Quality Factor of the Transient Evoked Oto-acoustic Emissions (TEOAE)

ENT 318 Artificial Organs Physiology of Ear

Auditory Physiology PSY 310 Greg Francis. Lecture 30. Organ of Corti

Comment by Delgutte and Anna. A. Dreyer (Eaton-Peabody Laboratory, Massachusetts Eye and Ear Infirmary, Boston, MA)

Chapter 11: Sound, The Auditory System, and Pitch Perception

Acoustics, signals & systems for audiology. Psychoacoustics of hearing impairment

Can You Hear Me Now?

The frequency analysis of the cochlea a review of Nobili et al (1998) and Ruggero et al (1992)

HEARING AND PSYCHOACOUSTICS

Structure, Energy Transmission and Function. Gross Anatomy. Structure, Function & Process. External Auditory Meatus or Canal (EAM, EAC) Outer Ear

Modeling auditory-nerve responses for high sound pressure levels in the normal and impaired auditory periphery

Signals, systems, acoustics and the ear. Week 5. The peripheral auditory system: The ear as a signal processor

Distortion product emissions from a cochlear model with nonlinear mechanoelectrical transduction in outer hair cells

The Structure and Function of the Auditory Nerve

Auditory Periphery! external middle inner. stapes movement initiates a pressure wave in cochlear fluid

SOLUTIONS Homework #3. Introduction to Engineering in Medicine and Biology ECEN 1001 Due Tues. 9/30/03

Measurement of cochlear power gain in the sensitive gerbil ear

SPHSC 462 HEARING DEVELOPMENT. Overview Review of Hearing Science Introduction

Pitfalls in behavioral estimates of basilar-membrane compression in humans a)

Cochlear anatomy, function and pathology II. Professor Dave Furness Keele University

Understanding Otoacoustic Emissions Generation

Auditory Physiology Richard M. Costanzo, Ph.D.

Consequences of Location-Dependent Organ of Corti Micro-Mechanics

Feed-Forward and Feed-Backward Amplification Model from Cochlear Cytoarchitecture: An Interspecies Comparison

Auditory System Feedback

Onset of basilar membrane non-linearity re ected in cubic distortion tone input-output functions

Educational Module Tympanometry. Germany D Germering

EMANATIONS FROM RESIDUUM OSCILLATIONS IN HUMAN AUDITORY SYSTEM

Intro to Audition & Hearing

Efferent-mediated control of basilar membrane motion

Issues faced by people with a Sensorineural Hearing Loss

Biological Basis of Hearing-Aid Design

Chapter 2 Auditory Periphery: From Pinna to Auditory Nerve

Stimulus Coding in the Auditory Nerve. Neural Coding and Perception of Sound 1

external middle inner

OtoAcoustic Emissions (OAE s)

Hearing Sound. The Human Auditory System. The Outer Ear. Music 170: The Ear

Music 170: The Ear. Tamara Smyth, Department of Music, University of California, San Diego (UCSD) November 17, 2016

COCHLEAR MODEL FOR NORMAL AND DAMAGED EARS

Lecture 3: Perception

Estimates of compression at low and high frequencies using masking additivity in normal and impaired ears

Systems Neuroscience Oct. 16, Auditory system. http:

A Bidirectional Analog VLSI Cochlear Model

Microstructures in the Organ of Corti Help Outer Hair Cells Form Traveling Waves along the Cochlear Coil

Frequency refers to how often something happens. Period refers to the time it takes something to happen.

Advanced otoacoustic emission detection techniques and clinical diagnostics applications

Cochlear compression in listeners with moderate sensorineural hearing loss

INTRODUCTION J. Acoust. Soc. Am. 102 (2), Pt. 1, August /97/102(2)/1038/11/$ Acoustical Society of America 1038

Basilar membrane nonlinearity determines auditory nerve rate-intensity functions and cochlear dynamic range

Receptors / physiology

A variant temporal-masking-curve method for inferring peripheral auditory compression a)

I. INTRODUCTION.

Auditory-based Automatic Speech Recognition

Modelling the micromechanics of the cochlea in Femlab

Cochlear anatomy, function and pathology I. Professor Dave Furness Keele University

Modeling of Mechanoelectrical Transduction of Hair Cells to Action Potentials in the Auditory Nerve

In vivo outer hair cell length changes expose the active process in the cochlea

INTEGRATING THE COCHLEA S COMPRESSIVE NONLINEARITY IN THE BAYESIAN APPROACH FOR SPEECH ENHANCEMENT

Nonlinear Cochlear Signal Processing

Detailed Simulation of the Cochlea: Recent Progress Using Large Shared Memory Parallel Computers

Application of force to the cochlear wall: effect on auditory thresholds, OHC transduction, and DPOAEs

Computational Models of Mammalian Hearing:

Temporal Adaptation. In a Silicon Auditory Nerve. John Lazzaro. CS Division UC Berkeley 571 Evans Hall Berkeley, CA

Human Acoustic Processing

Christopher J. Plack Department of Psychology, University of Essex, Wivenhoe Park, Colchester CO4 3SQ, England

What does it mean to analyze the frequency components of a sound? A spectrogram such as that shown here is the usual display of frequency components

Christopher J. Plack and Ray Meddis Department of Psychology, University of Essex, Wivenhoe Park, Colchester CO4 3SQ, United Kingdom

MECHANISM OF HEARING

Essential feature. Who are cochlear implants for? People with little or no hearing. substitute for faulty or missing inner hair

Sound and Hearing. Decibels. Frequency Coding & Localization 1. Everything is vibration. The universe is made of waves.

College of Medicine Dept. of Medical physics Physics of ear and hearing /CH

Limiting Frequency of the Cochlear Amplifier Based on Electromotility of Outer Hair Cells

FAST MOTILITY OF ISOLATED MAMMALIAN AUDITORY SENSORY CELLS. Hans P. Zenner, Ulrike Zimmermann and Alfred H. Gitter

Chapter 1: Introduction to digital audio

The mammalian cochlea possesses two classes of afferent neurons and two classes of efferent neurons.

Transcription:

Acoust. Sci. & Tech. 39, 1 (018) PAPER #018 The Acoustical Society of Japan Possible mechanisms of cochlear two-tone suppression represented by vector subtraction within a model Yasuki Murakami 1; and Shunsuke Ishimitsu ;y 1 National Institute of Technology, Oshima College, 1091 1 Komatsu, Suo-oshima Oshima, Yamaguchi, 74 193 Japan Graduate School of Information Science, Hiroshima City University, 3 4 1 Ozuka-Higashi, Asa-Minami, Hiroshima, 731 3194 Japan (Received 16 January 017, Accepted for publication 8 July 017) Abstract: This paper investigates possible mechanisms of cochlear two-tone suppression (TS) in models of the hair cell transducer and the cochlea. The hair cell transducer model can be represented by a saturation function. To simulate cochlear mechanics, a nonlinear transmission line cochlear model is used. The mechano-electric transducer curve of the outer hair cell (OHC) is regarded as the source of nonlinearity in cochlear mechanics. The saturation function approximated by a power series can explain TS in the OHC transducer model. However, this mathematical formulation cannot account for cochlear TS because the cochlear mechanics is more complicated than the saturation in the OHC transducer. To clarify two-tone interference graphically, it is expressed as a residual vector, the entries of which are the frequencies of the probe and suppressor. In this construct, the stronger of two tones introduced to the system nonlinearly reduces the output by vector subtraction. The model accomplished TS and displayed similar horizontal and vertical residual vectors. These analytical results suggest that TS is obtained from simple interference between the probe and the suppressor in the cochlear mechanics with nonlinear variation of the OHC transducer current. Keywords: Suppression, Cochlea, Hair cell, Model, Saturation PACS number: 43.64.Bt, 43.64.Jb, 43.64.Kc [doi:10.150/ast.39.11] 1. INTRODUCTION e-mail: murakami@oshima-k.ac.jp y e-mail: ishimitu@hiroshima-cu.ac.jp A distinctive feature of cochlear nonlinearity is twotone suppression (TS), in which one pure tone reduces the cochlear response to a second tone. The dominant and suppressed tones are called the suppressor and probe, respectively. This TS phenomenon, first observed in the auditory nerve (AN) half a century ago [1], has been detected in the basilar membrane (BM) [] as well as the inner hair cells (IHCs) [3] in the past few decades. Reference [] concluded that TS originates from mechanical phenomena at the BM and is generated by an active process. Some form of saturation is involved, as evidenced by such features of cochlear nonlinearity as the products of compression [4] and distortion [5]. According to experimental measurements, mechano-electric transduction by outer hair cells (OHCs) can be fitted by using the saturation function [6]. This function purportedly explains nonlinearities in cochlear mechanics [7]. To explain TS, the saturation function approximated by a power series shows that the output magnitude of one signal is suppressed when the input magnitude of the other signal is increased [8,9]. This mathematical formulation of TS is independent of the frequencies of the two signals with a constant output magnitude of the suppressed signal. Experimental measurement showed a frequency dependence of TS [10 1]. In these experiments, a lowfrequency suppressor generates the amplitude-modulated responses of the BM for a probe. For this phasic suppression, Geisler and Nuttall proposed a TS mechanism based on the saturation function imposed by a lowfrequency suppressor on the BM [10]. The low-frequency suppressor operates a rest point on the input output (IO) property of the saturation function. In contrast, the nearcharacteristic-frequency suppressor also produces TS on the BM response of the probe [,10 13]. Under this condition, the envelopes of temporal responses are constant [10,11]. Therefore, this tonic suppression cannot be explained by the rest-point operation on the saturation function as described in phasic suppression. These theo- 11

Acoust. Sci. & Tech. 39, 1 (018) retical and experimental studies indicate that the saturation function derived from the OHC transducer cannot fully explain both tonic and phasic suppression. The structure and dynamics of the cochlea are complicated. Vibration of the BM is generated by both fluid dynamics and mechanical motion in the cochlea. The classical cochlear transmission line model reveals the importance of both fluid dynamics and mechanical motion in reproducing BM motion for a tone [14,15]. From this standpoint, the cochlear TS phenomenon can be considered a consequence of both fluid dynamics and mechanical motion generated by two tones. Neely and Kim suggested that sharp tuning can be achieved by applying pressure to an active element on the BM within the transmission line model [16]. Other scholars have simulated TS using active elements that include simple saturation functions such as the hyperbolic tangent and the Boltzmann function [17 ]. These modeling studies demonstrate that TS is caused by a complicated interaction between the probe and the suppressor. However, as described previously, theoretical studies of simpler models such as the saturation function can account for TS without frequency dependence [8 10]. To explore the complicated mechanisms of TS in the cochlea, the interaction between the probe and the suppressor for TS should be addressed. However, understanding this interaction has remained elusive because TS has been separately analyzed for the probe and suppressor. In general, the IO properties for the probe and suppressor are widely used for the representation of TS [,10 13]. This representation can show a suppressed response caused by another tone; however, it cannot explain the influence of a tone on another tone. Instead of the IO property, a representation of the interference between the probe and the suppressor is required to investigate the influence of the saturating property in OHCs on cochlear TS. As mentioned earlier, transmission line models can reproduce TS. Consequently, these models can be used to explore TS mechanisms to help guide current experimental procedure. In this paper, we develop a cochlear transmission line model including an OHC model and propose a representation of suppression using vector subtraction to explain the relationship between the probe and the suppressor. In Sect., the cochlear mechanics, the models, and the concept of vector subtraction analysis are introduced. In Sect. 3, we show TS using the models. A discussion of the results is given in Sect. 4, followed by our conclusions in Sect. 5.. MODEL.1. Cochlear Mechanics The cochlea is shaped like a coiled duct and includes two chambers. The chambers are divided by the BM and Fig. 1 Fig. Vibration pattern of the BM in the cochlea. Scheme of cochlear mechanical model. are filled with fluid. An incoming sound wave affects the cochlear entrance and propagates in the cochlear fluid from the base to the apex. As a consequence of sound wave propagation, a pressure difference is generated between the two chambers. This pressure difference drives the mechanical vibration of the BM. Vibration patterns of the BM depend on the frequencies of the sound wave, as shown in Fig. 1. The IHCs and OHCs are located on the BM. The IHCs transform mechanical motion into neural information for the AN. In contrast, the OHCs sense mechanical vibration of the BM and exert pressure generated by its motilities [3,4] on the BM. Therefore, cochlear processing consists of feedback via the OHCs on the BM, as shown in Fig.. This OHC feedback amplifies the BM motion. However, the OHCs are not considered to be just an amplifier of the BM motion. It has been widely believed that the OHC feedback is the source of cochlear nonlinearities [7]... Hair Cell Transducer Hair cells consist of the soma and the hair bundle. BM vibration deflects the hair bundle of the hair cell. Displacement of the hair bundle opens ion channels located at the tip of the hair cell. The opening of the channels causes a voltage difference that drives the OHC motilities [3,4]. Lukashkin and Russell proposed a model for this hair cell mechano-electric transducer located on the tip of the hair cell [9]. The model shows TS of the signal processing in the hair cell. In this paper, we introduce the concept that the conductance of the hair cell mechano-electric transducer G tr ð c ðtþþ can be saturated by applying a secondorder Boltzmann function, which relates the probability of 1

Y. MURAKAMI and S. ISHIMITSU: VECTOR REPRESENTATION OF TWO-TONE SUPPRESSION 7 Transducer conductance Gtr [ns] 6 5 4 3 1 Ion Channels Hair Bundle Soma 0 00 100 0 100 00 300 400 Displacement ξ c [nm] Fig. 3 Conductance of the OHC transducer as a function of hair bundle displacement, fitted to Eq. (1) with parameter values listed in Table 1. Table 1 List of parameters in the OHC mechanoelectric transducer function (obtained from Ref. [9]). Parameter Value Unit G tr max 7 10 9 S a 1 65 10 6 m 1 a 16 10 6 m 1 c1 4 10 9 m c 41 10 9 m c rest :0 10 8 m a transducer channel opening to the displacement of a hair bundle. G tr max G tr ð c ðtþþ ¼ ½1 þ K ð1 þ K 1 ÞŠ ; ð1þ where t denotes time, G tr max is the maximal transducer conductance, and K 1 and K are variables given by K 1 ¼ exp½a 1 ð c1 c ðtþþ c rest ÞŠ; ðþ K ¼ exp½a ð c c ðtþþ c rest ÞŠ; where a 1, a, c1, and c are constants and c rest is the rest point of the hair bundle. Figure 3 shows the IO property of the hair cell transducer G tr in Eq. (1). Table 1 lists the parameter values of the transducer model..3. Cochlear Model Here, the one-dimensional transmission line model of the cochlea involving feedback from an active process [16] is expanded to include nonlinearity. The transmission line simulates the driving of the BM traveling wave using fluid dynamics. In the one-dimensional transmission line model, the traveling wave propagates in the x direction from the stapes to the helicotrema. P d denotes the pressure difference between the upper and lower scales of a box divided into compartments by the cochlear BM. P d drives the BM Fig. 4 Nonlinearized micromechanical model of the cochlea based on [16]. The constants for mass, damping, and stiffness are represented by m, c, and k, respectively. P d denotes the pressure in the fluid, P a is the pressure produced by the OHC, b indicates the BM displacement, and t is the TM displacement. displacement b. The macromechanical equation of the transmission line model is given by @ P d ðx; tþ ¼ @ b ðx; tþ ; ð3þ @x H @t where and H represent the fluid density and scale width, respectively. The boundary conditions at the basal and apical ends of the cochlea are given by @P d ðx; tþ @x ¼ s ðtþ; P d ðx; tþj x¼l ¼ 0; ð4þ x¼0 where s represents the inward displacement of the stapes footplate and the double-dot notation denotes its second time derivative. To produce the sharp tuning observed in the cat s AN, the BM and tectorial membrane (TM) are represented using the micromechanical model illustrated in Fig. 4. By denoting the pressure by Fðx; tþ ¼ðP d ðx; tþ P a ðx; tþ; 0Þ T and the displacement by ðx; tþ ¼ð b ðx; tþ; t ðx; tþþ T, the equations of motion of the micromechanical model can be written as where F ¼ M p @ ðx; tþ @ t þ C p @ðx; tþ @t M p ¼ m! 1 0 ; 0 m C p ¼ c 3 c þ c 3 c 1 þ c 3 c 3 K p ¼ k 3 k þ k 3 k 1 þ k 3 k 3 þ K p ðx; tþ; ð5þ! ; ð6þ! : 13

Acoust. Sci. & Tech. 39, 1 (018) The initial conditions are given by @ðx; tþ ðx; tþj t¼0 ¼ 0; @t ¼ 0; P a ðx; tþj t¼0 ¼ 0: ð7þ t¼0 The mechanical excitation of hair cells is assumed to result from the relative shearing displacement between the TM and the reticular lamina (RL). An active element senses the gap between the BM and the TM and provides feedback to the BM, thus amplifying its motion. We define the gap as the hair bundle displacement c, which depends on the location and time. c ðx; tþ ¼gðxÞ b ðx; tþ t ðx; tþ Here gðxþ is the lever gain between the BM displacement b and the radial displacement of the RL. The magnitude of the feedback P a is given by P a ðx; tþ ¼ðc 4 _ nl c ðx; tþþr 4 nl c ðx; tþþ; ð9þ where c 4 and r 4 are the damping and stiffness coefficients, respectively, and nl c is calculated from the mechanoelectric transducer G tr in Eq. (10). For the mechanoelectric transducer G tr, both the scale and offset of G tr ð c ðtþþ vary according to the following saturation function. nl c ðtþ ¼ tr½g tr ð c ðtþþ c rest Þ G tr ð c rest ÞŠ; ð10þ where c rest is the rest point of the hair bundle and tr is chosen such that nl c ðtþ ¼ cðtþ when the amplitudes of c ðtþ are <1 nm. For small displacements (<1 nm), our model reduces to Neely and Kim s model because nl c ¼ c in Eq. (10). The middle ear transmits ear drum vibrations driven by sound pressure P e to the cochlea. The middle ear is modeled as a mass spring damper system with one degree of freedom and the following equation of motion: P e ðtþ ¼m m s ðtþþc m _ s ðtþþk m s ðtþ; ð8þ ð11þ where m m, c m, and k m denote the mass, damping, and stiffness of the middle ear, respectively. The initial conditions are given by s ð0þ ¼0; _ s ð0þ ¼0: ð1þ.4. Vector Representation of Suppression We propose a graphical representation of suppression using vector subtraction. Suppression is conventionally calculated by subtracting the self-suppressing output shown in Fig. 5(a) (response to sinusoids with no temporal overlap) from the mutual suppressive output in Fig. 5(b) (response to simultaneous sinusoidal inputs). Self-suppression and TS are formulated as separate vectors. We also define the vector space set, in which the input frequencies of the probe and suppressor are represented as independent (a) Self-suppression (b) Mutual suppression (c) Concept of vector subtraction for TS Fig. 5 (a) Block diagram of the self-suppression process. (b) Block diagram of the TS process. (c) Vector representation of TS in the self-suppressed and twotone suppressed states, showing self-suppression and TS resulting from two temporally separated sinusoids and the simultaneous input of two sinusoids, respectively. The residual vector is a vector representing the difference between the two states. axes (Fig. 5(c)). Each element of the self-suppression vector is set by the responses of the saturation function to pure tones; the TS vector expresses the pair of tones entered into the saturation function. Vector subtraction then gives the difference between the self-suppression and TS. Engebretson and Eldredge mathematically formulated self-suppression, TS, and their difference [8]. The nonlinear saturation function GðÞ is expanded into a power series: GðÞ ¼tanhðÞ 3 3 : ð13þ By denoting the two sinusoidal inputs as 1 ¼ A 1 sin 1 and ¼ A sin, their outputs are calculated as 14

Y. MURAKAMI and S. ISHIMITSU: VECTOR REPRESENTATION OF TWO-TONE SUPPRESSION Gð 1 Þ A 1 A3 1 4 Gð Þ A A3 4 sin 1 þ distortion; sin þ distortion: ð14þ The distortions of these outputs contain the components 3 1 and 3, respectively. In their respective one-dimensional spaces 1 and, these outputs equal 1 and when A 1 and A are 1 and they saturate when A 1 and A are >1. When two sinusoids are simultaneously input as a TS, the output is Gð 1 þ Þ A 1 A3 1 sin 1 4 3A 1A þ A A3 4 3A 1 A þ distortions: sin ð15þ In this case, the distortions contain four components: þ 1, 1, 1 þ, and 1. The nonlinear output calculated by Eq. (15) clearly differs from that of self-suppression (such as the superposition of Eq. (14)). In Eq. (15), the saturations generated by the pair of sinusoids interfere with each other. The output can now be represented in a two-dimensional 1 space. To preserve the unique property of Eq. (15), the difference between the TS and self-suppression is expressed as Gð 1 þ Þ ðgð 1 ÞþGð ÞÞ 8 >< >: 3A 1 A 3A 1A sin if A 1 A, sin 1 if A 1 A. ð16þ Equation (16) indicates that the difference between the two outputs is reduced by the higher-amplitude sinusoid. Details of the calculation of both fundamentals and distortions are described in the Appendix. To graphically verify the interference between the two tones derived from the mathematical expression in Eq. (16), we propose a TS analysis method based on vector subtraction. Figure 6 highlights the differences between the vectors of the saturation function G in Eq. (13) for two arbitrary frequencies f 1 and f under the conditions of self-suppression shown in Fig. 5(a) and TS shown in Fig. 5(b). The vectors point vertically or horizontally when the input amplitude of one signal is greater than that of the other signal, which can be calculated by the saturation function given by Eq. (16)..5. Numerical Solution The transmission line model was solved in the time domain in two steps. In each time step, the boundary-value differential equation (Eq. (3)) and the initial-value differential equation (Eq. (5)) were solved in the first and second Fig. 6 Residual vectors in the saturation function G in Eq. (13) for inputs of two arbitrary frequencies. Table List of parameters in the micro-cochlear model, as described in [16]. Parameters of the middle-ear model were taken from [16]. Parameter Value Unit l 0.05 m N 500 1,000 kg/m 3 H 10 3 m W 10 3 m g 1 m 1 3 10 kg/m c 1 00 þ 1;500e 00x Ns/m 3 k 1 1:1 10 10 e 400x N/m 3 m 5 10 3 kg/m c 100e 0x Ns/m 3 k 7 10 7 e 440x N/m 3 c 3 0e 80x Ns/m 3 k 3 1 10 8 e 400x N/m 3 c 4 1:040 10 5 e 00x Ns/m 3 k 4 6:15 10 9 e 400x N/m 3 1 steps, respectively. The middle-ear model (Eq. (11)) was solved in the time domain given its initial condition (Eq. (1)). We selected the finite-difference method for the boundary-value problem and the Runge Kutta method for the initial-value problem, noting that Runge Kutta is the most commonly used numerical method in the context of the time domain [5]. The time step t was set to 3 ms. Table lists the parameter values of Neely and Kim s original model for simulating the cat cochlea. The BM tuning of mammalian cochlea is generally independent of the species [6]. For small peak displacements of the BM traveling wave (<1 nm), the gain in the cochlear amplifier at the base was set to 50 60 db, which is lower than specified in the Neely Kim model. High gains destabilize the dynamics of the Neely Kim model, even in the 15

Acoust. Sci. & Tech. 39, 1 (018) nonlinear case [7]. To reduce the high gain and resolve this instability problem, the damping coefficient in the cochlear model was computed as follows: :8e 10x c i 7! c i ; i ¼ 1; ; 3; 4: ð17þ Note that the increased damping does not affect the resonant frequency or the BM traveling wave distribution [7,8]. The frequency f p of the probe tone was 14 khz, and the frequencies f s of the suppressor tone were varied from 0.9 to 0 khz in 0.5 khz steps (excluding 14 khz). The intensities L p and L s were increased from 0 to 80 db in 10 db steps. As usual, the probe frequency f p was set to the characteristic frequency (CF) of the cochlear region (x ¼ 5:6 mm), defined as the frequency of maximum excitation at a 0 db input. Each of the two primary tones was simultaneously presented for 55 ms with rise/fall times of 5 ms. The model outputs were recorded for 10 ms, starting 40 ms after the tone presentation, to allow the system to reach a steady state. In the time domain, the model outputs were the BM velocity _ b ðx; tþ and BM displacement b ðx; tþ, which were separated into probe tone and suppressor tone components, respectively. In the frequency domain, the outputs were BM velocities _ b ðx; f p Þ and _ b ðx; f s Þ and BM displacements b ðx; f p Þ and b ðx; f s Þ. These components were obtained by fast Fourier transforms (FFTs) of their counterparts in the time domain. 3. RESULTS 3.1. Responses to a Single Tone Figure 7 (top) illustrates the amplitude of BM displacement obtained from the transmission line model at the CF as a function of the frequency and level of a single imposed tone with and without OHC involvement. In these plots, the CF was 14 khz, located at x ¼ 5:6 mm from the base. Note that the frequency response curves are sharper at lower input levels. The peak sharpness is characterized by the quality factor Q 10 (CF/bandwidth measured at 10 db below the peak). Q 10 was equal to 4.4 at 0 db, which is comparable to existing experimental data on animals [6]. The frequency responses broadened and the gain was reduced as the input intensity increased. At frequencies of less than half an octave below the CF, the OHC process had little effect on the BM displacement. However, close to the CF, the OHCs amplified the BM displacement by over 50 db relative to the case of no OHC involvement. To characterize the degree of compression in the frequency response curve, the rate of growth (ROG) has been defined as the ratio of the varying input level to the changed output level and was calculated at the various stimulus levels in the top panel of Fig. 7. The results with the OHC are presented in the bottom panel of Fig. 7. At frequencies of less than half an octave below the CF, the Fig. 7 (Top) Iso-intensity functions obtained from the transmission line model at CF ¼ 14 khz. The sound pressure was varied from 0 to 100 db in 0 db steps and the imposed frequency was varied from 0.3 to 0 khz. The solid and dashed lines represent the frequency responses of the complete model and the model excluding OHCs, respectively. (Bottom) ROG functions (degree of compression) obtained from the slopes of the frequency response curves. Linear and perfectly compressive growth occurred when ROG ¼ 1 db/db and ROG ¼ 0 db/db, respectively. ROGs closely approximated 1 db/db. The ROG values close to the CF were much smaller than 1 db/db (approximately 0.10 db/db), indicating compressive nonlinearity close to the CF. 3.. Two-tone Suppression in IO Functions Two-tone suppression has been observed at moderate sound pressure levels in db [,10 13]. To determine the input level of the OHC transducer model, we focused on the BM displacement obtained from the transmission line model shown in Fig. 7. In this case, the BM displacement varied from 10 to 1,000 nm at a moderate sound pressure level ranging from 0 to 80 db. Therefore, the input of the OHC transducer model was determined from 10 to 1,000 nm to simulate TS with the OHC transducer model. Figure 8 plots the IO function of the OHC transducer model (Eq. (1)). The left and right columns show the OHC transducer conductance by the probe and suppressor, respectively. Both the probe and suppressor displacements were increased from 10 nm (thin lines) to 1,000 nm (thick lines). The frequencies of the probe and suppressor were set to 14 and 13 khz, respectively. However, the IO property was not affected by the frequency as shown in 16

Y. MURAKAMI and S. ISHIMITSU: VECTOR REPRESENTATION OF TWO-TONE SUPPRESSION Fig. 8 IO function of the OHC transducer model in Eq. (1) for probe (left panel) and suppressor (right panel) as functions of displacement for both the probe and the suppressor. Displacements for the probe were varied from 10 nm (thinnest line) to 1,000 nm (thickest line) in 5 db steps. Eq. (16). The conductance for the probe decreased with increasing suppressor displacement. In the suppressor plots, the conductance decreased with increasing probe displacement. BM vibration is distributed spatially for two tones [9]. To avoid this problem, Murakami and Ishimitsu evaluated the energy of the BM vibration in terms of the overall response of the BM [9]. The energy of the BM vibration E b ð!þ is given by E b ð!þ ¼ 1 Z l Kð!Þ ð b n ð!þþ dx ð18þ where Kð!Þ ¼m 1 W! (! ¼ f ), f is the frequency of the sound, W is the width of the BM, and m 1 is the BM mass per unit area. Figure 9 plots the IO function of the BM energy at the CF produced by two-tone excitation as functions of the suppressor and probe levels. The numbers to the right represent frequency ratios. The left and right columns show the BM energy for the probe and suppressor, respectively. Both the probe and suppressor levels were increased from 0 db (thin lines) to 80 db (thick lines) at the four frequency ratios that were investigated ( f s =f p ¼ 0:11, 0.93, 1.07, and 1.43). For the low-frequency suppressor ( f s =f p ¼ 0:11), the BM energy for the probe was reduced at suppressor levels of 50 db and above. However, for the suppressor, the shapes of the IO functions did not change with the probe level. For similar frequencies ( f s =f p ¼ 0:93 and 1.07), the BM energy for the probe was constant at suppressor levels below 40 db. However, at suppressor levels of 40 db and above, the BM energy for the probe decreased with increasing suppressor level. In the suppressor plots, the BM energy for the probe decreased with increasing probe level. For the high-frequency suppressor ( f s =f p ¼ 1:43), the BM energy for the probe was reduced slightly. The BM energy for the suppressor decreased at probe levels of 70 db and above. 0 Fig. 9 IO function of BM energy for probe (left panels) and suppressor (right panels) as functions of the suppressor level L s and probe level L p. The probe tone was set to 14 khz with four suppressor tones. The intensities L s and L p were varied from 80 db (thickest line) to 0 db (thinnest line) in 10 db steps. Fig. 10 Residual vectors in the OHC transducer model in Eq. (1). Circles indicate the OHC response under the no-suppression condition (i.e., no temporal overlap between the probe and the suppressor). Arrows indicate shifts in the OHC response from the condition of no temporal overlap to the simultaneous input of the probe and suppressor. 3.3. Vector Representation of TS Figure 10 shows the differences in the vectors of the OHC transducer model in Eq. (1) under the conditions in Fig. 8. The difference vectors were calculated from the 17

Acoust. Sci. & Tech. 39, 1 (018) Fig. 11 Residual vectors in the cochlear transmission line model. Circles indicate the BM energy under the no-suppression condition (i.e., no temporal overlap between the probe and the suppressor). Arrows indicate shifts in the BM energy from the condition of no temporal overlap to the simultaneous input of the probe and suppressor. Fig. 1 Probability histograms of similarity cos of two vectors v ohc and v bm plotted in Figs. 10 and 11, respectively. The cosine similarity was calculated from Eq. (19). The input ranges of the saturation functions were 10 1,000 nm. The input intensities L p and L s varied from 0 to 80 db in 5 db steps. The number of compared vectors was 169 in each panel. vectors of self-suppression and TS as shown in Fig. 5. For low-amplitude displacement, the difference vector does not exist. However, with increasing input amplitude, the vectors point vertically or horizontally when the input amplitude of one displacement is greater than of the other displacement. Figure 11 plots the vector differences of the given cochlear model under the conditions of self-suppression and TS for two suppressor frequencies. At f p =f s ¼ 0:11 (top of Fig. 11), the energy of the probe is clearly suppressed by the higher energy of the suppressor, whereas the energy of the suppressor is minimally reduced (note the slight shift toward the left). In the cases of f p =f s ¼ 0:93 and 1.07, the energy is greatly suppressed by the companion tone. Finally, at f p =f s ¼ 1:43, the energy of the probe is suppressed by the suppressor except at very high probe energies, where the energy of the suppressor is reduced. Let us compare the two vectors v ohc and v bm plotted in Figs. 10 and 11. We define the similarity between the two vectors using the dot product: similarity ¼ cos ¼ v ohc v bm jv ohc jjv bm j : ð19þ The resulting similarity ranges from 1, indicating that they are exactly opposite, to 1, meaning that they are identical, with 0 indicating no correlation and values in between indicating an intermediate similarity or dissimilarity. Figure 1 presents a histogram of the similarities of the vectors v ohc and v bm for the four suppressor frequencies ( f s =f p ¼ 0:11, 0.93, 1.07, and 1.43) used in the previous figures. The similarities are concentrated around 1. This indicates that the vectors v ohc and v bm were highly correlated. In particular, for f s =f p ¼ 0:93 and 1.07, the probabilities were approximately 100%. However, for 18

Y. MURAKAMI and S. ISHIMITSU: VECTOR REPRESENTATION OF TWO-TONE SUPPRESSION f s =f p ¼ 0:11 and 1.43, these values were reduced to approximately 40 and 70%, respectively. 4. DISCUSSION 4.1. Analysis of TS In this paper, we have sought to reproduce TS within the models of the OHC transducer and the cochlea shown in Figs. 8 and 9, respectively. These phenomena were analyzed by using the concept of vector subtraction. Under these conditions, Fig. 1 showed that the residual vectors of the OHC transducer and the cochlea shown in Figs. 10 and 11 match. These analytical results indicate a basic feature of TS, i.e., a stronger tone suppresses the response of a weaker tone. The probe and suppressor are the weaker tone and stronger tone, respectively. Despite the complicated cochlear mechanics, these phenomena were obtained from models of the OHC transducer and the cochlea. In this section, we discuss the link between the OHC transducer and the cochlea. The experimental measurement showed constant and temporally varying envelopes in the suppressed responses [,10 13]. The theoretical consideration provided mechanisms for both types of suppression, with the saturation function representing the OHC transducer [8 10]. The saturation function can account for tonic suppression induced by the near-cf suppressor [8,9]. In contrast, for phasic suppression induced by the low-cf suppressor, the rest-point operation on the saturation function was proposed [10]. However, the link between the OHC transducer and the cochlea has been unclear. It has been proposed that an active OHC process amplifies the cochlear response, as shown in Fig. 4.16 of Ref. [30]. The active OHC process has been widely accounted for by many investigators (for a review, see Ref. [31]). In the present cochlear model, BM responses to a pure tone are amplified and depend on the input level, as shown in Fig. 7. This result is realistic because it implies that compression is solely caused by attenuated cochlear amplification, itself imposed by the saturation properties of the OHCs. According to the IO property of the OHC transducer model, the output of the OHCs was linear at low displacement levels and saturated at higher levels, which is consistent with compression [4]. Figures 8 and 9 showed TS of the IO functions of the OHC transducer current and BM motion, respectively. Their natures match the theoretical data of the OHC transducer current [9] and the experimental measurement of BM motion [1]. For the near-cf suppressor, the shapes of the IO properties obtained from both models were similar. However, for lower- and higher-cf suppressors, the shapes of the IO functions showed different trends. To investigate the interference between the probe and the suppressor for TS, Figs. 10 and 11 showed TS of the vector representations of the modeling results. In these figures, horizontal and vertical vectors indicate that the stronger tones suppress the responses for the weaker tones, and angled vectors indicate that the responses for each tone are equally suppressed when the two-tone levels are closed. These trends shown in both the OHC transducer model and the cochlear model are consistent with the mathematical explanation of TS in Refs. [8,9] (see the Appendix). In particular, as mentioned previously, the shapes of the IO functions in the two models differed for lower- and higher- CF suppressors, as shown in Figs. 8 and 9. However, for these conditions, the present analysis method reveals the basic nature of TS shown in Fig. 11. Figure 1 showed a quantitative comparison of the residual vectors in Figs. 10 and 11. For the near-cf suppressor, the vectors in the models are identical. This result implies that the interference of the two tones as a function of the OHC transducer current directly affects TS on the BM motion. However, for the lower- and higher-cf suppressors, half of the residual vectors were not correlated. This trend can be accounted for by the fact that the BM responses for the lower-frequency tones are not suppressed by the higher-frequency tones as shown in Fig. 11. 4.. Advantages and Disadvantages of Vector Subtraction Representation for TS In this paper, we have proposed a vector subtraction representation for cochlear TS and have successfully analyzed cochlear TS as considered in Sect. 4.1. In this section, we discuss the advantages and disadvantages of the proposed analysis method. To explain cochlear TS, IO functions have been widely used [9 1]. In this paper, Figs. 8 and 9 showed the IO functions representing cochlear TS. The IO functions can be easily calculated from each frequency component. For this reason, this method can be easily applied to both experimental and theoretical data when the frequency components are separated. As shown in Fig. 5, the proposed vector subtraction representation is calculated from each frequency component. This calculation method is similar to the calculation of the IO function. Therefore, as with the previous method, the proposed method can also be easily applied to both experimental and theoretical data. For modeling studies, cochlear models can be classified into transmission line models [17 ], phenomenological models of the cochlea [3,33], and simple models of cochlear partition [34,35]. The purposes of these models depend on the subject of the study. Despite these different starting points, these models incorporate a simple saturation property representing the OHC system and can account for cochlear TS. The similarity of this model construction to that of the proposed cochlear model suggests that the 19

Acoust. Sci. & Tech. 39, 1 (018) vector subtraction method can be used to analyze TS on the BM motion in cochlear models. The proposed analysis method is based on the separation of each frequency component. In this paper, we employed the FFT to separate the frequency components. This method can effectively separate frequency components of steady-state responses. However, the timevarying response affects the accuracy of the calculation result. In fact, experimental measurements show TS on the time-varying BM response [10,1]. Therefore, when applying the proposed method to a time-varying response, it is necessary to consider the calculation inaccuracy. 5. CONCLUSION To investigate possible mechanisms for TS within a cochlear model including an OHC model, we developed an analysis method based on vector subtraction with a distinctive residual response to a two-tone input. Both models showed a similar TS nature of the IO functions and similar vectors representing suppression. As a consequence of the vector analysis, the following possible mechanisms of TS were suggested. First, the frequency dependence of cochlear TS cannot be explained solely by the OHC transducer model. Second, TS is obtained from simple interference between the probe and the suppressor in the cochlear mechanics with a nonlinear variation of the OHC transducer current. REFERENCES [1] N. Kiang, Discharge Patterns of Single Fibers in the Cat s Auditory Nerve (MIT Press, Cambridge, Mass., 1966). [] M. A. Ruggero, L. Robles and N. C. Rich, Two-tone suppression in the basilar membrane of the cochlea: Mechanical basis of auditory-nerve rate suppression, J. Neurophysiol., 68, 1087 1099 (199). [3] P. M. Sellick and I. J. Russell, Two-tone suppression in cochlear hair cells, Hear. Res., 1, 7 36 (1976). [4] N. P. Cooper, Compression in the peripheral auditory system, in Compression, S. Bacon and R. R. Fay, Eds. (Springer, New York, 005), pp. 18 61. [5] J. L. Goldstein, Auditory nonlinearity, J. Acoust. Soc. Am., 41, 676 689 (1967). [6] A. J. Hudspeth and D. P. Corey, Sensitivity, polarity, and conductance change in the response of vertebrate hair cells to controlled mechanical stimuli, Proc. Natl. Acad. Sci. USA, 74, 407 411 (1977). [7] W. M. Roberts and M. A. Rutherford, Linear and nonlinear processing in hair cells, J. Exp. Biol., 11, 1775 1780 (008). [8] A. M. Engebretson and D. H. Eldredge, Model for the nonlinear characteristics of cochlear potentials, J. Acoust. Soc. Am., 44, 548 554 (1968). [9] A. N. Lukashkin and I. J. Russell, A descriptive model of the receptor potential nonlinearities generated by the hair cell mechanoelectrical transducer, J. Acoust. Soc. Am., 103, 973 980 (1998). [10] C. D. Geisler and A. L. Nuttall, Two-tone suppression of basilar membrane vibrations in the base of the guinea pig cochlea using low-side suppressors, J. Acoust. Soc. Am., 10, 430 440 (1997). [11] N. P. Cooper, Two-tone suppression in cochlear mechanics, J. Acoust. Soc. Am., 99, 3087 3098 (1996). [1] W. S. Rhode, Mutual suppression in the 6 khz region of sensitive chinchilla cochleae, J. Acoust. Soc. Am., 11, 805 818 (007). [13] A. L. Nuttall and D. F. Dolan, Two-tone suppression of inner hair cell and basilar membrane responses in the guinea pig, J. Acoust. Soc. Am., 93, 390 400 (1993). [14] L. Peterson and B. Bogert, A dynamical theory of the cochlea, J. Acoust. Soc. Am.,, 369 381 (1950). [15] J. J. Zwislocki, Theory of the acoustical action of the cochlea, J. Acoust. Soc. Am.,, 778 784 (1950). [16] S. T. Neely and D. O. Kim, A model for active elements in cochlear biomechanics, J. Acoust. Soc. Am., 79, 147 1480 (1986). [17] E. Zwicker, Suppression and (f 1 f )-difference tones in a nonlinear cochlear preprocessing model with active feedback, J. Acoust. Soc. Am., 80, 163 176 (1986). [18] L. Kanis and E. de Boer, Two-tone suppression in a locally active nonlinear model of the cochlea, J. Acoust. Soc. Am., 96, 156 165 (1994). [19] R. Nobili and F. Mammano, Biophysics of the cochlea. II: Stationary nonlinear phenomenology, J. Acoust. Soc. Am., 99, 44 55 (1996). [0] K. M. Lim and C. R. Steele, A three-dimensional nonlinear active cochlear model analyzed by the WKB-numeric method, Hear. Res., 170, 190 05 (00). [1] B. Epp, J. L. Verhey and M. Mauermann, Modeling cochlear dynamics: Interrelation between cochlear mechanics and psychoacoustics, J. Acoust. Soc. Am., 18, 1870 1883 (010). [] J. Meaud and K. Grosh, Effect of the attachment of the tectorial membrane on cochlear micromechanics and two-tone suppression, Biophys. J., 106, 1398 1405 (014). [3] W. E. Brownell, C. R. Bader, D. Bertrand and Y. De Ribaupierre, Evoked mechanical responses of isolated cochlear outer hair cells, Science, 7, 194 196 (1985). [4] H. J. Kennedy, A. C. Crawford and R. Fettiplace, Force generation by the mammalian hair bundle supports a role in cochlear amplification, Nature, 433, 880 883 (005). [5] R. J. Diependaal, H. Duifhuis, H. W. Hoogstraten and M. A. Viergever, Numerical methods for solving one-dimensional cochlear models in the time domain, J. Acoust. Soc. Am., 8, 1655 1666 (1987). [6] L. Robles and M. A. Ruggero, Mechanics of the mammalian cochlea, Physiol. Rev., 81, 1305 135 (001). [7] E. M. Ku and S. J. Elliott, Statistics of instabilities in a state space model of the human cochlea, J. Acoust. Soc. Am., 14, 1068 1079 (008). [8] M. Lamar, J. Xin and Y. Qi, Signal processing of acoustic signals in the time domain with an active nonlinear nonlocal cochlear model, Signal Process., 86, 360 374 (006). [9] Y. Murakami and S. Ishimitsu, Input level dependence of distortion products generated by saturating feedback in a cochlear model, Acoust. Sci. & Tech., 37, 1 9 (016). [30] R. Patuzzi, Cochlear micromechanics and macromechanics, in The Cochlea, P. Dallos, A. N. Popper and R. R. Fay, Eds. (Springer, New York, 1996), pp. 186 57. [31] P. Dallos, The active cochlea, J. Neurosci., 1, 4575 4585 (199). [3] R. Meddis, L. P. O Mard and E. A. Lopez-Poveda, A computational algorithm for computing nonlinear auditory frequency selectivity, J. Acoust. Soc. Am., 109, 85 861 (001). [33] X. Zhang, M. G. Heinz, I. C. Bruce and L. H. Carney, A phenomenological model for the responses of auditory-nerve 0

Y. MURAKAMI and S. ISHIMITSU: VECTOR REPRESENTATION OF TWO-TONE SUPPRESSION fibers: I. Nonlinear tuning with compression and suppression, J. Acoust. Soc. Am., 109, 648 670 (001). [34] F. Julicher, D. Andor and T. Duke, Physical basis of two-tone interference in hearing, Proc. Natl. Acad. Sci. USA, 98, 9080 9085 (001). [35] R. Szalai, A. Champneys, M. Homer, D. Ó. Maoiléidigh, H. Kennedy and N. Cooper, Comparison of nonlinear mammalian cochlear-partition models, J. Acoust. Soc. Am., 133, 33 336 (013). APPENDIX: FUNDAMENTALS AND DISTORTIONS IN THE CUBIC SYSTEM The fundamentals and distortions in the cubic system 3 are obtained from sinusoidal inputs. First, we consider the input of a sinusoid, A sin, to the cubic system 3. The output of the system for the sinusoid is In this case, the output contains the fundamental and the cubic distortion 3. Next, we calculate the output of the cubic system for two sinusoids, A 1 sin 1 þ A sin, as follows: ða 1 sin 1 þ A sin Þ 3 ¼ A3 1 4 þ 3A 1A 3A 1 A 4 3A 1A 4 sin 1 þ A3 4 þ 3A 1 A ðsinð þ 1 Þþsinð 1 ÞÞ ðsinð 1 þ Þþsinð 1 ÞÞ: sin ða:þ In this case, the output contains the two fundamentals 1 and and the four distortions þ 1, 1, 1 þ, and 1. ða sin Þ 3 ¼ A3 4 ðsin sin 3Þ: ða :1Þ 1