1: \documentclass[12pt]{elsart}
2: \usepackage{latexsym}
3: \usepackage{epsfig}
4: \usepackage{amsmath}
5:
6: %\input amssym.def
7: %\input amssym.tex
8: %\input epsf
9:
10: \newcommand{\be}{\begin{equation}}
11: \newcommand{\ee}{\end{equation}}
12: \newcommand{\br}{\begin{eqnarray}}
13: \newcommand{\er}{\end{eqnarray}}
14:
15: \newcommand{\TM}{T_{\mathrm{M}}}
16: \newcommand{\fm}{f_{\mathrm{m}}}
17: \newcommand{\zetam}{\zeta_{\mathrm{m}}}
18: \newcommand{\aM}{a_{\mathrm{M}}}
19: \newcommand{\Zm}{Z_{\mathrm{m}}}
20: \newcommand{\pe}{p_{\mathrm{e}}}
21: \newcommand{\ps}{p_{\mathrm{s}}}
22: \newcommand{\mm}{m_{\mathrm{m}}}
23: \newcommand{\cm}{c_{\mathrm{m}}}
24: \newcommand{\km}{k_{\mathrm{m}}}
25: \newcommand{\Mp}{M_{\mathrm{p}}}
26: \newcommand{\Cp}{C_{\mathrm{p}}}
27: \newcommand{\Kp}{K_{\mathrm{p}}}
28: \newcommand{\Ca}{C_{\mathrm{a}}}
29: \newcommand{\Ka}{K_{\mathrm{a}}}
30: \newcommand{\Lp}{L_{\mathrm{p}}}
31: \newcommand{\La}{L_{\mathrm{a}}^{n}}
32: \newcommand{\Mf}{M_{\mathrm{f}}}
33: \newcommand{\Mfs}{M_{\mathrm{f}}^{\mathrm{s}}}
34: \newcommand{\Ds}{D_{\mathrm{s}}}
35:
36: \newcommand{\mass}{\ \mathrm{g}\cdot\mathrm{cm}^{-2}}
37: \newcommand{\stif}{\ \mathrm{dyn}\cdot\mathrm{cm}^{-3}}
38: \newcommand{\damp}{\ \mathrm{dyn}\cdot\mathrm{s}\cdot\mathrm{cm}^{-3}}
39:
40: \def\R{{\Bbb R}}
41: \def\lap{\triangle}
42:
43: \renewcommand{\theequation}{\arabic{section}.\arabic{equation}}
44:
45: \begin{document}
46:
47: \begin{frontmatter}
48:
49: \title{Signal processing of acoustic signals in the time domain with an active nonlinear nonlocal
50: cochlear model}
51:
52: \author{M. Drew LaMar\thanksref{MDLemail}}
53: \address{Department of Mathematics, University of Texas at Austin,
54: Austin, TX 78712, USA.}
55: \thanks[MDLemail]{Corresponding author (mlamar@math.utexas.edu)}
56: \author{Jack Xin}
57: \address{Department of Mathematics and ICES (Institute of Computational
58: Engineering and Sciences), University of Texas at Austin, Austin, TX 78712, USA.}
59: \author{Yingyong Qi}
60: \address{Qualcomm Inc,
61: 5775 Morehouse Drive,
62: San Diego, CA 92121, USA.}
63:
64: \date{}
65: \setcounter{page}{1}
66: \setcounter{section}{0}
67: \begin{abstract}
68: A two space dimensional active nonlinear nonlocal cochlear model is
69: formulated in the time domain
70: to capture nonlinear hearing effects such as compression,
71: multi-tone suppression and difference tones. The
72: micromechanics of the basilar membrane (BM) are incorporated to model
73: active cochlear properties. An active gain parameter
74: is constructed in the form of
75: a nonlinear nonlocal functional of BM displacement.
76: The model is discretized with
77: a boundary integral method and
78: numerically solved using an iterative second order accurate
79: finite difference scheme.
80: A block matrix structure of the discrete
81: system is exploited to simplify the numerics with no loss
82: of accuracy. Model responses to multiple frequency stimuli
83: are shown in agreement with hearing experiments. A nonlinear spectrum
84: is computed from the model, and compared with FFT spectrum for noisy
85: tonal inputs.
86: The discretized model
87: is efficient and accurate, and can serve as a useful
88: auditory signal processing tool.
89: \end{abstract}
90:
91: \begin{keyword}
92: Auditory signal processing \sep cochlea \sep nonlinear filtering \sep basilar membrane \sep time domain
93: \end{keyword}
94:
95: \end{frontmatter}
96:
97: \newpage
98:
99: \section{Introduction}
100: \setcounter{equation}{0}
101: Auditory signal processing based on phenomenological models of
102: human perception has helped to
103: advance the modern technology of
104: audio compression \cite{pohlmann}.
105: It is of interest therefore to
106: develop a systematic mathematical framework
107: for sound signal processing based on models of the ear.
108: The biomechanics of
109: the inner ear (cochlea)
110: lend itself well to mathematical
111: formulation (\cite{allen,deboer} among others). Such
112: models can recover main aspects of the physiological data
113: \cite{bekesy,ruggero} for simple acoustic inputs (e.g. single frequency tones).
114: In this paper, we study a nonlinear nonlocal model and associated numerical
115: method for processing complex signals (clicks and noise) in the time domain.
116: We also obtain a new spectrum of sound signals with nonlinear hearing
117: characteristics which can be of potential interest for applications such
118: as speech recognition.
119:
120: Linear frequency domain cochlear models have been around for a long time and studied
121: extensively \cite{neely,neelykim}. The cochlea, however, is known to have
122: nonlinear characteristics, such as compression, two-tone suppression and combination tones, which
123: are all essential to capture interactions of multi-tone complexes \cite{dengg,dengk,jxintime}. In this
124: nonlinear regime, it is more expedient to work in the time domain to resolve complex nonlinear frequency
125: responses with sufficient accuracy. The nonlinearity in our model
126: resides in the outer hair cells (OHC's), which act as an amplifier to
127: boost basilar membrane (BM) responses to low-level
128: stimuli, so called active gain. It has been shown \cite{deboernut} that this type of nonlinearity is also nonlocal in
129: nature, encouraging near neighbors on the BM to interact.
130:
131: One space dimensional transmission line models with nonlocal nonlinearities have been studied previously
132: for auditory signal processing \cite{dengk,jxinglobal,jxintime,jxinPDE}.
133: Higher dimensional models
134: give sharper tuning curves and higher frequency selectivity.
135: In section 2, we begin with a two space dimensional (2-D)
136: macromechanical partial differential equation (PDE) model.
137: We couple the 2-D model with the BM micromechanics
138: of the active linear system in \cite{neelykim}.
139: We then make the gain parameter nonlinear
140: and nonlocal to complete the model setup, and
141: do analysis to simplify the model.
142:
143: In section 3, we discretize the system and formulate a second order accurate
144: finite difference scheme so as to combine efficiency and accuracy. The matrix we need to invert at
145: each time step has a time-independent part (passive) and a time-dependent part (active). In order
146: to speed up computations, we split the matrix into the passive
147: and active parts and devise an
148: iterative scheme. We only need to invert the passive part once, thereby significantly speeding
149: up computations. The structure of the system also allows us to reduce the complexity of the problem
150: by a factor of two, giving even more computational efficiency. A proof of convergence of the iterative scheme is
151: given in the Appendix.
152:
153: In section 4,
154: we discuss numerical results and show that
155: our model successfully reproduces the nonlinear effects such as compression,
156: multi-tone suppression, and combination difference tones. We demonstrate such effects by inputing
157: various signals into the model,
158: such as pure tones, clicks and noise. A nonlinear spectrum is
159: computed from the model and compared with FFT spectrum for the acoustic input of a single tone
160: plus Gaussian white noise. The conclusions are in section 5.
161:
162: \begin{figure}[tb]
163: \epsfig{file=Pictures/cochlea12.eps,width=350pt}
164: \caption{The figure on the left is a schematic of the cochlea, while the figure on the right
165: represents the upper chamber with the macromechanical equations and boundary conditions.}
166: \label{fig:coch}
167: \end{figure}
168:
169: \section{Model Setup}
170:
171: \subsection{Macromechanics}
172:
173: The cochlea consists of an upper and lower fluid filled chamber, the scala vestibuli and scala tympani,
174: with a shared elastic boundary called the basilar membrane (BM) (see Figure \ref{fig:coch}). The BM acts like a Fourier
175: transform with each location on the BM tuned to resonate at a particular frequency, ranging
176: from high frequency at the basal end to low frequency at the apical end. The acoustic wave
177: enters the ear canal, where it vibrates the eardrum and then is filtered through the middle ear,
178: transducing the wave from air to fluid in the cochlea via the stapes footplate. A traveling
179: wave of fluid moves from the base to the apex, creating a skew-symmetric motion of the BM.
180: The pressure difference drives the BM, which resonates according to the frequency content of the passing wave.
181:
182: We start with simplification of the upper cochlear chamber into a two dimensional rectangle
183: $\Omega = [0,L] \times [0,H]$ (see Figure \ref{fig:coch}). Due to the symmetry, we can ignore the lower
184: chamber. The bottom boundary ($z = 0$) is the BM, while the left boundary ($x = 0$) is the stapes
185: footplate. The macromechanical equations are
186: \begin{equation}
187: \left\{ \begin{array}{ll}
188: \varDelta p(x,z,t) = \frac{\partial^2 p}{\partial x^2} + \frac{\partial^2 p}{\partial z^2} = 0,\ x \in [0,L],\ z \in [0,H],\ t \in [0,\infty) \\
189: p_{x}(0,z,t) = \TM\pe(t),\ \ p(L,z,t) = 0 \\
190: p_{z}(x,0,t) = 2\rho u_{tt},\ \ p_{z}(x,H,t) = 0
191: \end{array}
192: \right.
193: \label{eq:macro}
194: \end{equation}
195: where $p(x,z,t)$ is the pressure difference across the BM, $u(x,t)$ denotes BM displacement, and
196: $\rho$ is fluid density.
197:
198: At the stapes footplate $(0,z)$, $\pe(t)$ is pressure at the eardrum while $\TM$ is a bounded
199: linear operator on the space of bounded continuous functions that incorporates the middle ear
200: filtering characteristics. In the frequency domain, for each input $\e^{\mathrm{i}\omega t}$,
201: $\TM(\omega) = 2\rho \mathrm{i}\omega/\Zm$, where $\Zm$ is the impedance of the middle ear.
202: The middle ear amplification function is given by $\aM = |\TM|$. In our case, based on Guinan and Peake
203: \cite{guinan},
204: \begin{equation}
205: \aM(f) = 1.815f^2((1-\frac{f^2}{\fm^{2}})^2 + (2\zetam f/\fm)^2)^{-1/2},
206: \label{eq:ssmear}
207: \end{equation}
208: where $\fm = 4\ \mathrm{kHz}$ is the middle ear characteristic frequency and $\zetam = 0.7$ is the middle
209: ear damping ratio. Thus, for $\pe(t) = A \exp\{2\pi \mathrm{i}f\} + c.c$, where c.c. is complex conjugate, we
210: have $\TM\pe(t) = B \exp\{2\pi \mathrm{i}f\} + c.c.$, where $B = \aM(f)A$.
211:
212: For more complex stimuli, it is useful to model the middle ear in the time domain as a one-degree of freedom
213: spring-mass system. The equivalent time domain formulation of the steady state middle ear is given by
214: \begin{equation}
215: \left\{ \begin{array}{l}
216: \pe(t) = \mm\ddot{s}(t) + \cm\dot{s}(t) + \km s(t) \\
217: s(0) = \dot{s}(0) = 0
218: \end{array}
219: \right.
220: \label{eq:tdmear}
221: \end{equation}
222: where $s(t)$ is stapes displacement and $\mm$, $\cm$ and $\km$ are the mass, damping and stiffness of the middle ear.
223: The stapes boundary condition in (\ref{eq:macro}) is replaced by
224: \begin{equation}
225: p_{x}(0,z,t) = 2\rho\ddot{s}(t)
226: \label{eq:bcmear}
227: \end{equation}
228: One of the interesting effects of using the time domain middle ear model
229: is that it reduces the dispersive instability in the
230: cochlea (see \cite{jxindisp}). It appears that
231: the steady state middle ear model ignores important transient effects and
232: phase shifts that help to reduce the shock to the cochlea.
233:
234: At the helicotrema $(L,z)$, we have used the
235: Dirichlet boundary condition $p(L,z,t) = 0$. In \cite{neelykim},
236: they used an absorbing boundary condition $p_{x}(L) = cp_{t}(L)$, where $c$ is a positive constant \cite{neelykim}.
237: Other models use the Neumann condition $p_{x}(L) = 0$. It has been stated that the frequency domain solutions
238: are minimally affected by which boundary condition is chosen \cite{neely}, and thus we have chosen the simpler Dirichlet condition.
239: However, interesting results on choosing the best initial conditions to minimize transient effects (dispersive
240: instability) has been shown in \cite{jxindisp} using the Neumann condition.
241: To summarize, the macromechanics consist of equations (\ref{eq:macro})--(\ref{eq:bcmear}).
242: \bigskip
243:
244: \subsection{Micromechanics}
245:
246: \begin{figure}[tb]
247: \parbox{2in}{\epsfig{file=Pictures/springmass.eps,width=100pt}}
248: \parbox{3in}{\caption{Cross section micromechanics of the cochlea. The mass $m_{1}$ represents a cross section
249: of the BM, while mass $m_{2}$ is a cross section of the TM. (Reconstructed from Figure 3 in \cite{neelykim})}\label{fig:sm}}
250: \end{figure}
251:
252: This is a resonant tectorial membrane model based on
253: \cite{neelykim}. The BM and TM
254: (tectorial membrane) are modeled
255: as two lumped masses coupled by a spring and damper, with each mass connected to a wall
256: by a spring and damper. (See Figure \ref{fig:sm}). A classical approximation
257: is to have no longitudinal coupling except that which occurs through the fluid. Denoting $\xi(x,t) = (u(x,t),v(x,t))$
258: as BM and TM displacement, respectively, the equations of motion for the passive case at each point along the
259: cochlea are given by
260: \begin{equation}
261: \Mp\ddot{\xi} + \Cp\dot{\xi} + \Kp\xi = F
262: \label{eq:passive}
263: \end{equation}
264: where
265: \begin{equation}
266: \Mp = \left[\begin{array}{cc}
267: m_{1} & 0 \\
268: 0 & m_{2}
269: \end{array}
270: \right], \ \
271: \Cp = \left[\begin{array}{cc}
272: c_{1}+c_{3} & -c_{3} \\
273: -c_{3} & c_{2}+c_{3}
274: \end{array}
275: \right], \ \
276: \Kp = \left[\begin{array}{cc}
277: k_{1}+k_{3} & -k_{3} \\
278: -k_{3} & k_{2}+k_{3}
279: \end{array}
280: \right]
281: \label{eq:passmat}
282: \end{equation}
283: and forcing function
284: \begin{equation}
285: F = \left[\begin{array}{c}
286: p(x,0,t) \\
287: 0
288: \end{array}
289: \right]
290: \label{eq:force}
291: \end{equation}
292: The parameters $m_{i}$, $c_{i}$, and $k_{i}$ are functions of $x$. The initial conditions are given by
293: \begin{equation}
294: \xi(x,0) = \dot{\xi}(x,0) = 0
295: \label{eq:ic}
296: \end{equation}
297: To make the model active, a self-excited vibrational force acting on the BM is added to (\ref{eq:passive}):
298: \[
299: \Mp\ddot{\xi} + \Cp\dot{\xi} + \Kp\xi = F + F_{\mathrm{a}}
300: \]
301: where
302: \[
303: F_{\mathrm{a}} = \left[\begin{array}{c}
304: \gamma[c_{4}(\dot{u} - \dot{v}) + k_{4}(u-v)] \\
305: 0
306: \end{array}
307: \right]
308: \]
309: The difference $u-v$ represents OHC displacement.
310: The parameter $\gamma \in [0,1]$ is the active gain control. In \cite{neelykim}, this is a constant, but in our
311: case will be a nonlinear nonlocal functional of BM displacement and BM location. Bringing $F_{\mathrm{a}}$ to the left, we have
312: \begin{equation}
313: \Mp\ddot{\xi} + (\Cp-\gamma \Ca)\dot{\xi} + (\Kp-\gamma \Ka)\xi = F
314: \label{eq:active}
315: \end{equation}
316: where
317: \begin{equation}
318: \Ca = \left[\begin{array}{cc}
319: c_{4} & -c_{4} \\
320: 0 & 0
321: \end{array}
322: \right], \ \
323: \Ka = \left[\begin{array}{cc}
324: k_{4} & -k_{4} \\
325: 0 & 0
326: \end{array}
327: \right]
328: \label{eq:actmat}
329: \end{equation}
330: Thus, the micromechanics consist of equations (\ref{eq:passmat})--(\ref{eq:actmat}).
331:
332: \subsection{Nonlinear Nonlocal Active Gain}
333: A compressive nonlinearity in the model is necessary to capture effects such as two-tone suppression and combination
334: tones. Also, to allow for smoother BM profiles, we make the active gain nonlocal. Thus we have
335: \[
336: \hat{u}(x,t) = \frac{2}{\sqrt{\lambda\pi}}\int_{0}^{L}\e^{-(x-s)^2/\lambda}u^{2}(s,t)\ \d s
337: \]
338: and gain
339: \[
340: \gamma(x,t) = \frac{1}{1+\theta\hat{u}}
341: \]
342: where $\theta,\ \lambda$ are constants.
343:
344: \subsection{Semi-discrete Formulation}
345: Solving the pressure Laplace equation on the rectangle using separation of variables, we arrive at
346: \begin{equation}
347: p(x,0,t) = \TM\pe(t)(x-L) + \sum_{n=1}^{\infty}A_{n}\cos\beta_{n}x
348: \label{eq:sov}
349: \end{equation}
350: where
351: \begin{equation}
352: A_{n} = \left(\frac{-4\rho H}{L}\right)\left(\frac{\coth\beta_{n}H}{\beta_{n}H}\right)\int_{0}^{L}u_{tt}(x,t)\cos\beta_{n}x\d x
353: \label{eq:integral}
354: \end{equation}
355: \[\beta_{n} = \frac{(n-\frac{1}{2})\pi}{L}\]
356: Substituting (\ref{eq:sov}) into (\ref{eq:force}) and then discretizing (\ref{eq:active}) in space into $N$ grid points, we have
357: \begin{equation}
358: M\vec{\xi}_{tt} + C(t)\vec{\xi}_{t} + K(t)\vec{\xi} = \vec{b}(t)
359: \label{eq:disc}
360: \end{equation}
361: where
362: \[M = \left[\begin{array}{cc}
363: M_{1}+\alpha \Mf & 0 \\
364: 0 & M_{2}
365: \end{array}
366: \right]
367: \]
368: \[C(t) = \Cp - \hat{\Gamma}(t) \Ca = \left[\begin{array}{cc}
369: C_{1}+C_{3}-\Gamma(t) C_{4} & -(C_{3}-\Gamma(t) C_{4}) \\
370: -C_{3} & C_{2}+C_{3}
371: \end{array}
372: \right]
373: \]
374: \[K(t) = \Kp - \hat{\Gamma}(t) \Ka = \left[\begin{array}{cc}
375: K_{1}+K_{3}-\Gamma(t) K_{4} & -(K_{3}-\Gamma(t) K_{4}) \\
376: -K_{3} & K_{2}+K_{3}
377: \end{array}
378: \right]
379: \]
380: \[\vec{b}(t) = \left[\begin{array}{c}
381: \TM\pe(t)(\vec{x}-L) \\
382: 0
383: \end{array}
384: \right]
385: \]
386: \[M_{\mathrm{f},ij} = \sum_{k=1}^{K}\frac{\coth\beta_{k}H}{\beta_{k}H}\cos(\beta_{k}x_{i})\cos(\beta_{k}x_{j})w_{j}\]
387: \[\alpha = \frac{4\rho H}{N-1}\]
388: $\Cp$, $\Kp$, $\Ca$ and $\Ka$ are now block diagonal, where $K_{i} = \hbox{diag}\{k_{i}\}$ and $C_{i} = \hbox{diag}\{c_{i}\}$.
389: Also, $M_{i} = \hbox{diag}\{m_{i}\}$, $\Gamma(t) = \hbox{diag}\{\gamma_{i}(t)\}$ and $\hat{\Gamma}(t) = \hbox{diag}\{\Gamma(t),0\}$.
390: The numbers $w_{j}$ are numerical integration weights in the discretization of (\ref{eq:integral}) and are chosen based on the
391: desired degree of accuracy.
392: Note that we can write $\Mf = \Mfs W$, where $W = \hbox{diag}(w_{j})$
393: and $\Mfs$ is symmetric and positive definite. The result of separation of variables produced the matrix $\Mf$,
394: which is essentially the mass of fluid on the BM and {\em dynamically couples} the system.
395:
396: \section{Numerics}
397: In formulating a numerical method, we note that the matrices in (\ref{eq:disc}) can be split into a time-independent passive
398: part and a time-dependent active part. In splitting in this way, we are able to formulate an iterative scheme where we only
399: need to do one matrix inversion on the passive part for the
400: entire simulation. Thus, using second order approximations of the
401: first and second derivates in (\ref{eq:disc}), we arrive at
402: \begin{equation}
403: (\Lp-\La)\vec{\xi}^{n+1} = \vec{B}^{n} \;\;\; \Longrightarrow \;\;\; \vec{\xi}^{n+1,k+1} = \Lp^{-1}\vec{B}^{n} + \Lp^{-1}\La\vec{\xi}^{n+1,k}
404: \label{eq:iteration}
405: \end{equation}
406: where superscript $n$ denotes discrete time, $k$ denotes iteration and
407: \[\Lp = 2M + \frac{3}{2}\Delta{t}\Cp + \Delta{t}^{2}\Kp\]
408: \[\La = \hat{\Gamma}^{n}[\frac{3}{2}\Delta{t}\Ca + \Delta{t}^{2}\Ka]\]
409: \[\vec{B}^{n} = \Delta{t}^{2}\vec{b}(n\Delta{t}) +
410: M(5\vec{\xi}^{n} - 4\vec{\xi}^{n-1} + \vec{\xi}^{n-2}) +
411: \frac{\Delta{t}}{2}C^{n}(4\vec{\xi}^{n} - \vec{\xi}^{n-1}).\]
412: Proof of convergence will follow naturally from the next discussion. Notice that this is a
413: $2N \times 2N$ system. We shall simplify it to an $N \times N$ system
414: and increase the computational efficiency.
415:
416: \subsection{System Reduction}
417:
418: We write $\Lp$ and $\La$ in block matrix form as
419: \[\Lp = \left[\begin{array}{cc}
420: \tilde{M}_{1} & -P_{3} \\
421: -P_{3} & \tilde{M}_{2}
422: \end{array}
423: \right]
424: \]
425: \[\La = \left[\begin{array}{cc}
426: \Gamma^{n} P_{4} & -\Gamma^{n} P_{4} \\
427: 0 & 0
428: \end{array}
429: \right]
430: \]
431: where
432: \[\tilde{M}_{1} = 2(\alpha \Mf + M_{1}) + P_{1} + P_{3}\]
433: \[\tilde{M}_{2} = 2M_{2} + P_{2} + P_{3}\]
434: \[P_{i} = \frac{3}{2}\Delta{t}C_{i} + \Delta{t}^{2}K_{i}\]
435: It is easily seen that the left inverse of $\Lp$ is given by
436: \[
437: \Lp^{-1} = \left[\begin{array}{cc}
438: D^{-1} & D^{-1}\tilde{M}_{2}^{-1}P_{3} \\
439: \tilde{M}_{2}^{-1}P_{3}D^{-1} & \tilde{M}_{2}^{-1}P_{3}D^{-1}\tilde{M}_{1}P_{3}^{-1}
440: \end{array}
441: \right]
442: \]
443: where
444: \begin{eqnarray}
445: D & = & \tilde{M}_{1} - P_{3}\tilde{M}_{2}^{-1}P_{3} \nonumber \\
446: & = & 2\alpha \Mf + [2M_{1} + P_{1} + P_{3}(I - \tilde{M}_{2}^{-1}P_{3})] \nonumber \\
447: & = & \{2\alpha \Mfs + [2M_{1} + P_{1} + P_{3}(I - \tilde{M}_{2}^{-1}P_{3})]W^{-1}\}W \nonumber \\
448: & \equiv & \Ds W
449: \label{eq:dsdef}
450: \end{eqnarray}
451: Note that $D$ is invertible since $\Mfs$ is positive definite, thus invertible, and all other terms are positive diagonal
452: matrices, and thus their sum is positive definite and invertible. We also have
453: \begin{equation}
454: \Lp^{-1}\La = \left[\begin{array}{cc}
455: D^{-1}\Gamma^{n} P_{4} & -D^{-1}\Gamma^{n} P_{4} \\
456: \tilde{M}_{2}^{-1}P_{3}D^{-1}\Gamma^{n} P_{4} & -\tilde{M}_{2}^{-1}P_{3}D^{-1}\Gamma^{n} P_{4}
457: \end{array}
458: \right]
459: \label{eq:lpinvla}
460: \end{equation}
461: Letting $\vec{B}^{n} = (\vec{B}_{1}^{n},\vec{B}_{2}^{n})$, we have
462: \begin{eqnarray}
463: \vec{u}^{n+1,k+1} & = & W^{-1}[\zeta_{1}^{n} + \Ds^{-1}\Gamma^{n} P_{4}(\vec{u}-\vec{v})^{n+1,k}] \label{eq:iter1} \\
464: \vec{v}^{n+1,k+1} & = & W^{-1}\tilde{M}_{2}^{-1}P_{3}[\zeta_{2}^{n} + \Ds^{-1}\Gamma^{n} P_{4}(\vec{u}-\vec{v})^{n+1,k}] \label{eq:iter2}
465: \end{eqnarray}
466: where
467: \begin{equation}
468: \zeta_{1}^{n} = \Ds^{-1}[\vec{B}_{1}^{n} + \tilde{M}_{2}^{-1}P_{3}\vec{B}_{2}^{n}] \label{eq:zeta1}
469: \end{equation}
470: \begin{equation}
471: \zeta_{2}^{n} = \Ds^{-1}[\vec{B}_{1}^{n} + \tilde{M}_{1}P_{3}^{-1}\vec{B}_{2}^{n}] \label{eq:zeta2}
472: \end{equation}
473: At each time step, we do 2 $N \times N$ matrix solves in (\ref{eq:zeta1}) and (\ref{eq:zeta2}) to initialize the
474: iterative scheme. Then, since the same term appears in both equations (\ref{eq:iter1}) and (\ref{eq:iter2}), for
475: each $k$ we only have to do 1 $N \times N$ matrix solve. In practice, since $\Ds$ is symmetric,
476: positive definite and {\em time-independent}, we compute the Cholesky factorization of $\Ds$ at the start of the simulation
477: and use the factorization for more efficient matrix solves at each step. As a side note, if we subtract (\ref{eq:iter2}) from
478: (\ref{eq:iter1}), we have one equation for the OHC displacement $u-v$.
479:
480: \section{Numerical Results}
481:
482: \subsection{Model Parameters}
483: We start with a modification of the parameters in \cite{neelykim} (See Table \ref{tab:params}).
484: It is known that higher dimensional models give higher sensitivity. This is the case with this model. The 1-D
485: model \cite{neelykim} gives a 90 dB active gain at 16 kHz, whereas the 2-D model gives a 160 dB active gain. Thus, we need to
486: tune the system to reduce the gain. There are many ways to do this, and the method we choose is to increase all the damping
487: coefficients in the table by the following:
488: \[2\e^{0.2773x}c_{i} \mapsto c_{i},\ \ i=1,2,3,4\]
489:
490: \begin{table}[tb]
491: \footnotesize
492: \caption{Model parameters in cgs units}
493: \label{tab:params}
494: \begin{center}
495: \begin{tabular}{||l|l||l|l||} \hline
496: $m_{1}(x)$ & $3\cdot10^{-3}\mass$ & $\mm$ & $34.4\cdot10^{-3}\mass$ \\
497: $c_{1}(x)$ & $20+1500\e^{-2x}\damp$ & $\cm$ & $1.21\cdot10^{3}\damp$ \\
498: $k_{1}(x)$ & $1.1\cdot10^{9}\e^{-4x}\stif$ & $\km$ & $2.18\cdot10^{7}\stif$ \\
499: $m_{2}(x)$ & $0.5\cdot10^{-3}\mass$ & $L$ & $2.5$ cm \\
500: $c_{2}(x)$ & $10\e^{-2.2x}\damp$ & $H$ & $0.1$ cm \\
501: $k_{2}(x)$ & $7\cdot10^{6}\e^{-4.4x}\stif$ & $\rho$ & $0.1 \ \mathrm{g}\cdot\mathrm{cm}^{-3}$ \\
502: $c_{3}(x)$ & $2\e^{-0.8x}\damp$ & $\theta$ & $0.5$ \\
503: $k_{3}(x)$ & $10^{7}\e^{-4x}\stif$ & $\lambda$ & $0.08$ cm \\
504: $c_{4}(x)$ & $1040\e^{-2x}\damp$ & $\Delta t$ & $2.5\cdot10^{-6}$ -- $10^{-5}$ s\\
505: $k_{4}(x)$ & $6.15\cdot10^{8}\e^{-4x}\stif$ & $N$ & $401$ \\ \hline
506: \end{tabular}
507: \end{center}
508: \end{table}
509:
510: \subsection{Isointensity Curves}
511:
512: \begin{figure}[tb]
513: \centerline{\epsfig{file=Pictures/sens.eps,width=330pt,height=200pt}}
514: \caption{Both figures are sensitivity curves for CP = 0.77 cm or CF = 10 kHz. The left plot is a collection of
515: sensitivity curves for the linear steady state active model where the parameter is the active gain $\gamma$. The right
516: plot is a collection of sensitivity curves for the nonlinear time domain model where the parameter is pressure
517: at the eardrum in dB SPL (sound pressure level).}
518: \label{fig:sens}
519: \end{figure}
520:
521: In an isocontour plot, a probe is placed at a specific location on the BM where the time response is measured and analyzed for
522: input tones covering a range of frequencies. Figure \ref{fig:sens} shows isointensity curves for $\mathrm{CF} = 10\ \mathrm{kHz}$,
523: which corresponds to $\mathrm{CP} = 0.77\ \mathrm{cm}$. The characteristic place (CP) for a frequency is defined as the location on
524: the BM of maximal response from a pure tone of that frequency in the fully linear active model ($\gamma = 1$). The characteristic
525: frequency (CF) at a BM location is the inverse of this map. The left plot is the linear steady state active case. The parameter
526: is the active gain $\gamma$, and for each value of the active gain we get a curve that is a function of the input frequency.
527: The value of this function is the ratio $|u|(\mathrm{CP})/P_{\mathrm{e}}$, where $|u|(\mathrm{CP})$ is BM displacement at the
528: characteristic place and $P_{\mathrm{e}}$ is pressure at the eardrum. This is known as sensitivity. It is basically an output/input
529: ratio and gives the transfer characteristics of the ear at that particular active level. Notice that when $\gamma = 1$, the BM
530: at the characteristic place is most sensitive at the corresponding characteristic frequency, but at lower values of the gain, the
531: sensitivity peak shifts to lower frequencies.
532:
533: Analogously, the second plot in Figure \ref{fig:sens} shows isointensity curves for the nonlinear time domain model where now the
534: parameter is the intensity of the input stimulus in dB SPL (sound pressure level). For the time domain, we measure the root-mean-square BM amplitude from
535: 5 ms (to remove transients) up to a certain time $T$. Note that for high-intensity tones, the model becomes passive while
536: low-intensity tones give a more active model. This shows {\it compression}. Again, there is a frequency shift of the sensitivity
537: peak (about one-half octave) from low to high-intensity stimuli in agreement with \cite{ruggero}, so called half-octave shift. The
538: plot agrees well with Figure 5 in \cite{ruggero}.
539:
540: \subsection{Complex Stimuli}
541:
542: \begin{figure}[tb]
543: \centerline{\epsfig{file=Pictures/click_final.eps,width=330pt,height=200pt}}
544: \caption{An impulse, or click, lasting 0.1 ms starting at 0.4 ms is input into the nonlinear nonlocal ear model. The left plot is the BM displacement
545: time series for various CF's ranging from 0.5-4 kHz. The right plot is a sensitivity plot for various stimulus intensities
546: at CF = 6.4 kHz.}
547: \label{fig:click}
548: \end{figure}
549:
550: The first non-sinusoidal input we look at is a click. In the experiment in the left plot of Figure \ref{fig:click},
551: we put probes at varying characteristic places associated with frequencies ranging from 0.5-4 kHz to measure the time
552: series BM displacement. The click was 40 dB with duration 0.1 ms starting at 0.4 ms. All responses were normalized to
553: amplitude 1. The plot is similar to Figure 4 in \cite{dengk}. In the right plot of Figure \ref{fig:click}, a probe
554: was placed at CP for 6.4 kHz and the time series BM volume velocity was recorded for various intensities and the
555: sensitivity plotted. This shows, similar to Figure \ref{fig:sens}, the compression effects at higher intensities. See
556: Figure 9 in \cite{ruggero} for a similar plot.
557:
558: \begin{figure}[tb]
559: \centerline{\epsfig{file=Pictures/noise_final.eps,width=330pt,height=200pt}}
560: \caption{Gaussian noise is input into the ear. The left plot is the BM displacement times series for various CF's ranging
561: from 0.5-16 kHz. The right plot is a sensitivity plot for CF = 6.4 kHz.}
562: \label{fig:noise}
563: \end{figure}
564:
565: The second non-sinusoidal input we explore is Gaussian white noise. Figure \ref{fig:noise} is similar in all regards
566: to Figure \ref{fig:click}. Notice again in the right plot the compression effect.
567:
568: \bigskip
569:
570: \subsection{Difference Tones}
571:
572: \begin{figure}[tb]
573: \centerline{\epsfig{file=Pictures/cdt_final.eps,width=330pt,height=200pt}}
574: \caption{Two sinusoidal tones, 7 and 10 kHz at 80 dB each, are the input. The left and middle plots are snapshots at 15 ms of BM
575: displacement and active gain, respectively. The right plot is a spectrum plot of the BM displacement time series at CP for 4 kHz.}
576: \label{fig:cdt}
577: \end{figure}
578:
579: Any nonlinear system with multiple sinusoidal inputs will create difference tones. If
580: two frequencies $f_{1}$ and $f_{2}$ are put into the ear, $nf_{1} \pm mf_{2}$ will be created at varying intensities,
581: where $n$ and $m$ are nonnegative integers. The cubic difference tone, denoted $f = 2f_{1}-f_{2}$, where
582: $f_{1} < f_{2}$, is the most prominent. Figure \ref{fig:cdt} contains three plots of one experiment. The experiment
583: consists of two sinusoidal tones, 7 and 10 kHz at 80 dB each. The cubic difference tone is 4 kHz. The plot on the
584: left is the BM profile for the experiment at 15 ms. We see combination tone peaks at 1.21 cm (CP for 4 kHz),
585: 1.54 cm (CP for 2 kHz) and 1.85 cm (CP for 1 kHz). The middle plot shows the snapshot at 15 ms of the active gain parameter,
586: showing the difference tones getting an active boost. Finally, the right plot is a spectrum plot of the time series for BM
587: displacement at 1.21 cm, the characteristic place for 4 kHz. The cubic difference tone is above 1 nm and can therefore
588: be heard.
589:
590: \subsection{Multi-tone Suppression}
591:
592: \begin{figure}[tb]
593: \centerline{\epsfig{file=Pictures/supp.eps,width=330pt,height=200pt}}
594: \caption{Isodisplacement curves at CP for 4 kHz showing responses to single tones (dashed line w/ circle) and responses to
595: the same tones in the presence of high-side and low-side suppressors presented at 80 dB SPL.}
596: \label{fig:supp}
597: \end{figure}
598:
599: \begin{figure}[tb]
600: \epsfig{file=Pictures/multisupp.eps,width=380pt}
601: \caption{Spectrum plots of BM responses for characteristic frequencies along the BM, from 500 Hz to 16 kHz, with 50 dB noise
602: and a 2 kHz tone ranging from 40-80 dB. R0 is the average of the BM response spectrum of 0 dB noise from 0.5-16 kHz. The solid
603: line represents noise with tone, the dotted line noise without tone.}
604: \label{fig:multisupp}
605: \end{figure}
606:
607: \begin{figure}[tb]
608: \epsfig{file=Pictures/multisuppfft.eps,width=380pt}
609: \caption{Spectrum plots of input signals consisting of 50 dB noise and a 2 kHz tone ranging from 40-80 dB. N0 is the average
610: of the spectrum of 0 dB noise from 0.5-16 kHz. The solid line represents noise with tone, the dotted line noise without tone.}
611: \label{fig:multisuppfft}
612: \end{figure}
613:
614: Two-tone (and multi-tone) suppression is characteristic of a compressive nonlinearity and has been recognized in the
615: ear \cite{ruggero,dengg,geisler}. Figure \ref{fig:supp} illustrates two-tone suppression and is a collection of isodisplacement
616: curves that show decreased tuning in the presence of suppressors and is similar to Figure 16 in \cite{ruggero}. We placed
617: a probe at the CP for 4 kHz (1.21 cm) and input sinusoids of various frequencies. At each frequency, we record the pressure at
618: the eardrum that gives a 1 nm displacement for 4 kHz in the FFT spectrum of the time series response at CP. The curve
619: without suppressors is dashed with circles. We then input each frequency again, but this time in the presence of a low side
620: (0.5 kHz) tone and high side (7.5 kHz) tone, both at 80 dB. Notice the reduced tuning at the CF. Also notice the asymmetry of
621: suppression, which shows low side is more suppressive than high side, in agreement with \cite{geisler}.
622:
623: For multi-tone suppression, we look at tonal suppression of noise. In Figure \ref{fig:multisupp},
624: for each plot, a probe was placed at every grid point along the BM and the time response was measured from 15 ms up to 25 ms.
625: The signal in each consisted of noise at 50 dB with a 2 kHz tone ranging from 40 dB to 80 dB (top to bottom).
626: An FFT was performed for each response and its characteristic frequency amplitude was recorded and plotted in decibels
627: relative to the average of the response spectrum of 0 dB noise from 0.5-16 kHz. We see suppression of all frequencies, with again low-side suppression stronger
628: than high-side suppression. Figure \ref{fig:multisupp} is qualitatively similar to Figure 3 in \cite{dengg}. It is useful to
629: compare this figure with Figure \ref{fig:multisuppfft}. This figure is the same as Figure \ref{fig:multisupp}, except we
630: do an FFT of the input signal at the eardrum. Comparing these two figures shows that we have a new spectral transform that can
631: be used in place of an FFT in certain applications, for example
632: signal recognition and noise suppression.
633:
634: \section{Conclusions}
635:
636: We studied a two-dimensional nonlinear nonlocal variation of the
637: linear active model in \cite{neelykim}.
638: We then developed an efficient and accurate numerical method and used this method to explore nonlinear effects
639: of multi-tone sinusoidal inputs, as well as clicks and noise. We showed numerical results illustrating compression,
640: multi-tone suppression and difference tones.
641: The model reached agreement with experiments \cite{ruggero} and
642: produced a novel nonlinear spectrum. In future work, we will analyze the model
643: responses to speech and resulting spectra for speech
644: recognition.
645: It is also interesting to study
646: the inverse problem \cite{sondhi}
647: of finding efficient and automated ways
648: to tune the model to different physiological data.
649: Applying the model to
650: psychoacoustic signal processing \cite{jxinPDE} will be another
651: fruitful line of inquiry.
652: \vspace{.1 in}
653:
654: \begin{ack}
655: The work was partially supported by NSF grant ITR-0219004.
656: J. X. would like to acknowledge a fellowship from the John Simon Guggenheim Memorial Foundation, and a
657: Faculty Research Assignment Award at UT Austin.
658: \end{ack}
659:
660: \begin{appendix}
661:
662: \section{Appendix: Convergence of Iterative Scheme (\ref{eq:iteration})}
663:
664: We need the following Lemma:
665: \begin{lem}
666: If
667: \[M = \left[ \begin{array}{cc}
668: A & -A \\
669: B & -B
670: \end{array}
671: \right]
672: \]
673: then every non-zero eigenvalue of $M$ is an eigenvalue of $A-B$.
674: \end{lem}
675: \begin{pf}
676:
677: Let $\lambda$ be a non-zero eigenvalue of $M$ with non-trivial eigenvector $\vec{x} = (\vec{x}_{1},\vec{x}_{2})$. Thus,
678: $M\vec{x} = \lambda\vec{x}$ gives
679: \begin{equation}
680: A(\vec{x}_{1}-\vec{x}_{2}) = \lambda\vec{x}_{1} \label{equ:A1}
681: \end{equation}
682: \begin{equation}
683: B(\vec{x}_{1}-\vec{x}_{2}) = \lambda\vec{x}_{2} \label{equ:A2}
684: \end{equation}
685: Subtracting the two equations, we have
686: \[(A-B)(\vec{x}_{1}-\vec{x}_{2}) = \lambda(\vec{x}_{1}-\vec{x}_{2})\]
687: Now, if $\vec{x}_{1}-\vec{x}_{2} = 0$, then from \ref{equ:A1} and \ref{equ:A2} above and $\lambda \neq 0$, we have
688: $\vec{x}_{1} = \vec{x}_{2} = 0$. But this means $\vec{x} = 0$, which is a contradiction. Thus, $\lambda$ is an
689: eigenvalue of $A-B$ with non-trivial eigenvector $\vec{x}_{1}-\vec{x}_{2}.$\qed
690: \end{pf}
691:
692: \begin{thm}
693: There exists a constant $C > 0$ such that if $\Delta{t} < C$, then
694: \[\rho(\Lp^{-1}\La) < 1\]
695: where $\rho$ is the spectral radius. Thus, the iterative scheme converges.
696: \end{thm}
697: \begin{pf}
698:
699: By the above lemma applied to (\ref{eq:lpinvla}), with constant $\gamma$, we have
700: \begin{eqnarray*}
701: \sigma(\Lp^{-1}\La) & \subset & \gamma\sigma(D^{-1}P_{4}-\tilde{M}_{2}^{-1}P_{3}D^{-1}P_{4}) \\
702: & = & \gamma\sigma[(I - \tilde{M}_{2}^{-1}P_{3})D^{-1}P_{4}]
703: \end{eqnarray*}
704: where $\sigma$ denotes spectrum. Thus, we have
705: \begin{eqnarray*}
706: \rho(\Lp^{-1}\La) & \leq & \gamma||(I-\tilde{M}_{2}^{-1}P_{3})W^{-1}\Ds^{-1}P_{4}||_{2} \\
707: & \leq & \gamma||(I-\tilde{M}_{2}^{-1}P_{3})W^{-1}||_{2}||\Ds^{-1}||_{2}||P_{4}||_{2}
708: \end{eqnarray*}
709: Now, let $(\lambda,\vec{x})$ be the eigen-pair of $\Ds$ with $\lambda$ the smallest eigenvalue and $||\vec{x}|| = 1$.
710: Note that $\lambda > 0$ since $\Ds$ is positive definite. Thus, we have $1/\lambda$ is the largest eigenvalue of $\Ds^{-1}$,
711: which gives
712: \[||\Ds^{-1}||_{2} \leq 1/\lambda\]
713: Thus, using the definition of $\Ds$ from (\ref{eq:dsdef}), we have
714: \begin{eqnarray*}
715: \lambda & = & \vec{x}^{T}\Ds\vec{x}\\
716: & = & \vec{x}^{T}\{2\alpha \Mfs + [2M_{1}+P_{1}+P_{3}(I-\tilde{M}_{2}^{-1}P_{3})]W^{-1}\}\vec{x}\\
717: & \geq & \vec{x}^{T}\{[2M_{1}+P_{1}+P_{3}(I-\tilde{M}_{2}^{-1}P_{3})]W^{-1}\}\vec{x}\\
718: & \geq & \min\{[2m_{1}+p_{1}+p_{3}(1-\tilde{m}_{2}^{-1}p_{3})]w^{-1}\}
719: \end{eqnarray*}
720: where lowercase represents diagonal entries. The third line above follows from $2\alpha \Mfs$ being positive definite.
721: Finally, we have
722: \begin{eqnarray*}
723: \rho(\Lp^{-1}\La) & \leq & \gamma||(I-\tilde{M}_{2}^{-1}P_{3})W^{-1}||_{2}||\Ds^{-1}||_{2}||P_{4}||_{2} \\
724: & \leq & \gamma\frac{\max[(1-\tilde{m}_{2}^{-1}p_{3})w^{-1}]\max(p_{4})}{\min\{[2m_{1}+p_{1}+p_{3}(1-\tilde{m}_{2}^{-1}p_{3})]w^{-1}\}}
725: \end{eqnarray*}
726: For $\Delta{t}$ small enough, we have convergence.\qed
727: \end{pf}
728:
729: With our parameters, for convergence it is {\it sufficient} that $\Delta{t} \leq 0.0008$. In practice, however, convergence
730: is seen for $\Delta{t}$ as large as $0.01$.
731: \end{appendix}
732:
733: \begin{thebibliography}{99}
734:
735: \bibitem{allen}J. B. Allen,
736: ``Cochlear Modeling-1980'',
737: in: M. Holmes and L. Rubenfeld, eds.,
738: Lecture Notes in Biomathematics, Springer-Verlag,
739: 1980, Vol. 43, pp. 1--8
740:
741: \bibitem{deboer}E. de Boer,
742: ``Mechanics of the Cochlea: Modeling Efforts'',
743: in: P. Dollas, A. Popper and R. Fay,
744: Springer Handbook of Auditory Research, Springer-Verlag,
745: 1996, pp. 258--317
746:
747: \bibitem{deboernut}E. de Boer and A. L. Nuttall,
748: ``Properties of Amplifying Elements in the Cochlea'',
749: in: A. W. Gummer, ed.,
750: Biophysics of the Cochlea: From Molecules to Models,
751: Proc. Internat. Symp.,
752: Titisee, Germany, 2002
753:
754: \bibitem{dengg}L. Deng and C. D. Geisler,
755: ``Responses of auditory-nerve fibers to multiple-tone complexes'',
756: J. Acoust. Soc. Amer.,
757: Vol. 82, No. 6, 1987, pp. 1989--2000
758:
759: \bibitem{dengk}L. Deng and I. Kheirallah,
760: ``Numerical property and efficient solution of a transmission-line model for basilar membrane wave motions'',
761: Signal Processing,
762: Vol. 33, 1993, pp. 269--285
763:
764: \bibitem{geisler}C. D. Geisler,
765: From Sound to Synapse,
766: Oxford University Press, Oxford, 1998
767:
768: \bibitem{guinan}J. J. Guinan and W. T. Peake,
769: ``Middle-ear characteristics of anesthesized cats'',
770: J. Acoust. Soc. Amer.,
771: Vol. 41, No. 5, 1967, pp. 1237--1261
772:
773: \bibitem{neely}S. T. Neely,
774: ``Mathematical modeling of cochlear mechanics'',
775: J. Acoust. Soc. Amer.,
776: Vol. 78, No. 1, July 1985, pp. 345--352
777:
778: \bibitem{neelykim}S. T. Neely and D. O. Kim,
779: ``A model for active elements in cochlear biomechanics'',
780: J. Acoust. Soc. Amer.,
781: Vol. 79, No. 5, May 1986, pp. 1472--1480
782:
783: \bibitem{pohlmann}K. Pohlmann,
784: Principles of Digital Audio, 4th Edition,
785: McGraw-Hill Video/Audio Professional, 2000
786:
787: \bibitem{ruggero}L. Robles and M. A. Ruggero,
788: ``Mechanics of the mammalian cochlea'',
789: Physiological Reviews,
790: Vol. 81, No. 3, July 2001, pp. 1305--1352
791:
792: \bibitem{sondhi}M. Sondhi,
793: ``The Acoustical Inverse Problem for the Cochlea'',
794: in: M. Holmes and L. Rubenfeld, eds.,
795: Lecture Notes in Biomathematics,
796: Springer-Verlag, 1980, pp. 95--104
797:
798: \bibitem{bekesy}G. von B\'{e}k\'{e}sy,
799: Experiments in Hearing, McGraw-Hill, New York, 1960
800:
801: \bibitem{jxindisp}J. Xin,
802: ``Dispersive instability and its minimization in time domain computation of steady state responses of cochlear models'',
803: J. Acoust. Soc. Amer.,
804: Vol. 115, No. 5, Pt. 1, 2004, pp. 2173--2177
805:
806: \bibitem{jxinPDE}J. Xin and Y. Qi,
807: ``A PDE based two level model of the masking property of the human ear'',
808: Comm. Math. Sci., Vol. 1, No. 4, 2003, pp. 833-840
809:
810: \bibitem{jxinglobal}J. Xin and Y. Qi,
811: ``Global well-posedness and multi-tone solutions of a class of nonlinear nonlocal cochlear models in hearing'',
812: Nonlinearity,
813: Vol. 17, 2004, pp. 711--728
814:
815: \bibitem{jxintime}J. Xin, Y. Qi, and L. Deng,
816: ``Time domain computation of a nonlinear nonlocal cochlear model with applications to multitone interaction in hearing'',
817: Comm. Math. Sci.,
818: Vol. 1, No. 2, 2003, pp. 211--227
819:
820: \end{thebibliography}
821:
822: \end{document}
823: