PROPAGATION OF SOUND WAVES IN DUCTS



Similar documents
Physics 9e/Cutnell. correlated to the. College Board AP Physics 1 Course Objectives

Introduction to acoustic imaging

3.0 Exhaust Muffler Design Principles. 3.1 Basic Concepts

6 J - vector electric current density (A/m2 )

Finite Element Analysis for Acoustic Behavior of a Refrigeration Compressor

Sound Power Measurement

SOUND INTENSITY AND ITS MEASUREMENT AND APPLICATIONS

A wave lab inside a coaxial cable

High Speed Aerodynamics Prof. K. P. Sinhamahapatra Department of Aerospace Engineering Indian Institute of Technology, Kharagpur

Dynamic Process Modeling. Process Dynamics and Control

18 Q0 a speed of 45.0 m/s away from a moving car. If the car is 8 Q0 moving towards the ambulance with a speed of 15.0 m/s, what Q0 frequency does a

Differential Relations for Fluid Flow. Acceleration field of a fluid. The differential equation of mass conservation

Section 5.0 : Horn Physics. By Martin J. King, 6/29/08 Copyright 2008 by Martin J. King. All Rights Reserved.

Statistical Energy Analysis software

Dispersion diagrams of a water-loaded cylindrical shell obtained from the structural and acoustic responses of the sensor array along the shell

State Newton's second law of motion for a particle, defining carefully each term used.

SCHWEITZER ENGINEERING LABORATORIES, COMERCIAL LTDA.

Sound absorption and acoustic surface impedance

Transmission Line and Back Loaded Horn Physics

Acoustic Terms, Definitions and General Information

Acoustics: the study of sound waves

Relationship between Sound Pressure and Sound Power Levels

PAGE 2. Figure 1: Difference between PWL ins and SPL 1m

BIOMEDICAL ULTRASOUND

T = 1 f. Phase. Measure of relative position in time within a single period of a signal For a periodic signal f(t), phase is fractional part t p

Introduction to Engineering System Dynamics

INTERNATIONAL JOURNAL OF RESEARCH SCIENCE & MANAGEMENT

Notes on Elastic and Inelastic Collisions

Lesson 11. Luis Anchordoqui. Physics 168. Tuesday, December 8, 15

Basic Concepts of Sound. Contents: Definitions db Conversion Sound Fields db ± db

State Newton's second law of motion for a particle, defining carefully each term used.

Waves - Transverse and Longitudinal Waves

3-D WAVEGUIDE MODELING AND SIMULATION USING SBFEM

7. DYNAMIC LIGHT SCATTERING 7.1 First order temporal autocorrelation function.

Sound Attenuation INTRODUCTION

Chapter 2. Derivation of the Equations of Open Channel Flow. 2.1 General Considerations

Development and optimization of a hybrid passive/active liner for flow duct applications

AS COMPETITION PAPER 2008

RANDOM VIBRATION AN OVERVIEW by Barry Controls, Hopkinton, MA

1) The time for one cycle of a periodic process is called the A) wavelength. B) period. C) frequency. D) amplitude.

Integration of a fin experiment into the undergraduate heat transfer laboratory

CO MPa (abs) 20 C

Section 2.0 : Construction and Measurement of a Simple Test Transmission Line

An octave bandwidth dipole antenna

Absorption Coefficients and Impedance Daniel A. Russell Science and Mathematics Department, Kettering University, Flint, MI, 48504

APPLICATION NOTE AP050830

TCOM 370 NOTES 99-4 BANDWIDTH, FREQUENCY RESPONSE, AND CAPACITY OF COMMUNICATION LINKS

Thnkwell s Homeschool Precalculus Course Lesson Plan: 36 weeks

1. Fluids Mechanics and Fluid Properties. 1.1 Objectives of this section. 1.2 Fluids

Sound. References: L.D. Landau & E.M. Lifshitz: Fluid Mechanics, Chapter VIII F. Shu: The Physics of Astrophysics, Vol. 2, Gas Dynamics, Chapter 8

Waves: Recording Sound Waves and Sound Wave Interference (Teacher s Guide)

ACOUSTIC CAVITIES DESIGN PROCEDURES

Chapter 10. Flow Rate. Flow Rate. Flow Measurements. The velocity of the flow is described at any

Exhaust noise control case study for 2800 class locomotive

Waves and Sound. AP Physics B

du u U 0 U dy y b 0 b

Linear Parameter Measurement (LPM)

Infrared Spectroscopy: Theory

Absorption mufflers in exhaust systems. Rolf Jebasinski. J. Eberspächer GmbH & Co. Abstract

Contents. Microfluidics - Jens Ducrée Physics: Fluid Dynamics 1

Chapter 17: Change of Phase

MODULE VII LARGE BODY WAVE DIFFRACTION

Lecture 6 - Boundary Conditions. Applied Computational Fluid Dynamics

SECTION 2 Transmission Line Theory

8.2 Elastic Strain Energy

An equivalent circuit of a loop antenna.

A LAMINAR FLOW ELEMENT WITH A LINEAR PRESSURE DROP VERSUS VOLUMETRIC FLOW ASME Fluids Engineering Division Summer Meeting

Fluid Mechanics Prof. S. K. Som Department of Mechanical Engineering Indian Institute of Technology, Kharagpur

NUMERICAL ANALYSIS OF THE EFFECTS OF WIND ON BUILDING STRUCTURES

WAVES AND FIELDS IN INHOMOGENEOUS MEDIA

State of Stress at Point

Antenna Properties and their impact on Wireless System Performance. Dr. Steven R. Best. Cushcraft Corporation 48 Perimeter Road Manchester, NH 03013

Module 1 : Conduction. Lecture 5 : 1D conduction example problems. 2D conduction

APPLICATION NOTE ULTRASONIC CERAMIC TRANSDUCERS

4.4 WAVE CHARACTERISTICS 4.5 WAVE PROPERTIES HW/Study Packet

XI / PHYSICS FLUIDS IN MOTION 11/PA

INTRODUCTION TO FLUID MECHANICS

Indiana's Academic Standards 2010 ICP Indiana's Academic Standards 2016 ICP. map) that describe the relationship acceleration, velocity and distance.

Assessment Plan for Learning Outcomes for BA/BS in Physics

Doppler Effect Plug-in in Music Production and Engineering

Fundamentals of Acoustics & Practical HVAC Design Considerations. Demir Doken Acoustic Engineer

Experiment 7: Familiarization with the Network Analyzer

Sample Questions for the AP Physics 1 Exam

Global Seasonal Phase Lag between Solar Heating and Surface Temperature

Unsteady Pressure Measurements

The Fourier Analysis Tool in Microsoft Excel

potential in the centre of the sphere with respect to infinity.

4 SENSORS. Example. A force of 1 N is exerted on a PZT5A disc of diameter 10 mm and thickness 1 mm. The resulting mechanical stress is:

How To Control Noise In A C7 Data Center Server Room

Estimating Acoustic Performance of a Cell Phone Speaker Using Abaqus

Chapter 22: Electric Flux and Gauss s Law

Keysight Technologies Understanding the Fundamental Principles of Vector Network Analysis. Application Note

Eðlisfræði 2, vor 2007

PART VIII: ABSORPTIVE SILENCER DESIGN

Physical Science Study Guide Unit 7 Wave properties and behaviors, electromagnetic spectrum, Doppler Effect

Fluids and Solids: Fundamentals

Thompson/Ocean 420/Winter 2005 Tide Dynamics 1

Audio Engineering Society. Convention Paper. Presented at the 119th Convention 2005 October 7 10 New York, New York USA

Rock Bolt Condition Monitoring Using Ultrasonic Guided Waves

The Physics of Music: Brass Instruments. James Bernhard

Transcription:

PROPAGATION OF SOUND WAVES IN DUCTS Finn Jacobsen Acoustic Technology, Department of Electrical Engineering Technical University of Denmark, Building 352 Ørsteds Plads, DK-2800 Lyngby Denmark fja@elektro.dtu.dk Technical University of Denmark September 2011 Note no 31260

2

CONTENTS page 1 Introduction... 5 2 Fundamentals of acoustic wave motion... 5 3 Plane waves in duct with rigid walls... 9 3.1 The sound field in a duct terminated by an arbitrary impedance... 11 3.2 Radiation of sound from an open-ended tube... 18 4 Sound transmission through coupled pipes... 22 4.1 The transmission matrix... 23 4.2 System performance... 28 4.3 Dissipative silencers... 33 5 Sound propagation in ducts with mean flow... 35 6 Three-dimensional waves in ducts with rigid walls... 36 6.1 The sound field in a duct with rectangular cross section... 37 6.2 The sound field in a duct with circular cross section... 44 6.3 The sound field in a duct with arbitrary cross-sectional shape... 51 7 Sound propagation in ducts with walls of finite impedance... 58 7.1 Ducts with nearly hard walls... 59 7.2 Lined ducts... 60 References... 63 Bibliography... 64 Appendix: Cylindrical Bessel and Neumann functions of integer order... 65 List of symbols... 67 Index... 69 3

4

1. INTRODUCTION Duct acoustics is the branch of acoustics concerned with sound propagation in pipes and ducts. Sound propagation within ducts is of practical concern in many areas of acoustics and noise control. Cylindrical cavities are used for testing or calibrating transducers, and impedance tubes are used for measuring the acoustic properties of absorbing materials. In a number of musical instruments (brass and woodwind instruments, organs) the sound is generated in pipes. Air-distributing systems for heating, ventilating and air-conditioning also transmit noise. Exhaust noise from petrol- and diesel-driven reciprocating engines must be reduced by silencers, which essentially are combinations of pipes. The purpose of this note is to introduce the topic. Non-linear effects are not dealt with, and the effects of viscosity, heat conduction and mean flow are only touched upon. 2. FUNDAMENTALS OF ACOUSTIC WAVE MOTION Sound waves in a fluid are oscillatory disturbances, generated, for example, by a vibrating surface. Such disturbances manifest themselves by pressure fluctuations, which is what we can hear, but also by similar perturbations of the temperature and the density, and by particles 1 moving back and forth. Conservation of mass implies that where u is the particle velocity (a vector), and (2.1) (2.2) is the density of the medium, which is the sum of the equilibrium value D 0 and the small, time-varying perturbation D. Under normal ambient conditions (101.3 kpa, 20 C) the density of air is 1.204 kgm -3. If eq. (2.1) can be approximated as follows: (2.3) (2.4) In a gas such as air sound is an adiabatic phenomenon, which means that there is no local heat exchange. 2 Under such conditions (2.5) 1 In acoustics the notion of particles and particle velocity refers to a macroscopic average of the medium, not to molecules. 2 Very near solid walls, for instance in narrow tubes, heat conduction cannot be ignored, and the process tends to be isothermal, which means that the temperature is constant. See chapter 7. 5

where (2.6) is the total pressure, which is the sum of the static (barometric) pressure p 0 and the acoustic, time-varying pressure p (henceforth called the sound pressure), K is a constant, and (2.7) is the ratio of specific heats ( 1.401 for air). Differentiating eq. (2.5) with respect to time gives (2.8) where we have introduced the quantity (2.9) which, as we shall see later, is the square of the speed of sound ( 343 ms -1 for air at 20 C). Combining eqs. (2.4) and (2.8) gives a relation between the divergence of the particle velocity and the sound pressure: (2.10) The quantity (2.11) is the adiabatic bulk modulus, which is the reciprocal of the adiabatic compressibility. Euler s equation of motion for a fluid is a relation between the gradient of the sound pressure and the acceleration of fluid particles, where 6 (2.12) (2.13) in which > is the particle displacement (or position) and the second term on the right side is due to the fact that the particle velocity is a function of time and position. This is a higherorder term. Thus we can simplify to (2.14)

The approximations of eqs. (2.4), (2.8) and (2.14) imply neglecting second- and higher-order terms in the small perturbations p and D. These equations, usually termed the linearised acoustic equations, are good approximations provided that (2.15) which, with a static pressure of 101.3 kpa, is seen to be the case even at a sound pressure level of 140 db re 20 :Pa (200 Pa). The linearisation simplifies the analysis enormously. 3 Adiabatic compression The absence of local heat exchange implies that the entropy of the medium per unit mass is constant [1], and that pressure fluctuations are accompanied by density fluctuations and temperature fluctuations. From eq. (2.5) it follows that which shows that or indicating that the fractional pressure increase associated with adiabatic compression because of the increase of the temperature exceeds the fractional increase of the density by a factor of (. The temperature fluctuations can be found in a similar manner from the ideal gas equation, where R is the gas constant ( 287 J@ kg -1 K -1 ) and T is the absolute temperature. It follows that and Taking the divergence of eq. (2.14) and combining with eq. (2.10) gives the linearised wave equation expressed in terms of the sound pressure: The Laplacian, that is, the operator L 2, takes the form (2.16) (2.17a) 3 Nonlinear effects include waveform distortion, interaction of waves (instead of linear superposition), and formation of shock waves. It is not only the level that matters; nonlinear effects accumulate with distance. See, e.g., ref. [2]. 7

in a Cartesian coordinate system. In a spherical coordinate system it is (2.17b) and in a cylindrical coordinate system it can be written (2.17c) Many acoustic phenomena, resonances in tubes and cavities, for example, can with advantage be studied in the frequency domain. Linearity implies that a disturbance that varies sinusoidally with time with the frequency T, a vibrating solid surface, for example, will generate a sound field in which the sound pressure p varies sinusoidally with time with the frequency T at all positions, and so will the particle velocity and the perturbations of the density and the temperature. It follows that we can describe any linear (first-order) quantity in such a harmonic sound field as the real part of a complex amplitude that varies with the position multiplied with a factor of e jtt, which takes account of the time dependence. This leads to the usual complex representation; see, e.g., ref. [3]. The sound pressure is now written as a complex amplitude multiplied with a complex exponential, and the real, physical sound pressure is the real part, (2.18) where n is the phase of the complex sound pressure at t = 0. Throughout this note complex harmonic variables are indicated by carets. With a time dependence of e jtt the operator M/Mt can be replaced by jt, and the operator M 2 /Mt 2 can be replaced by -T 2. The wave equation now becomes the Helmholtz equation, (2.19a) which is simpler to solve than equation (2.16). The Helmholtz equation is usually written in the form where (2.19b) (2.20) is the wavenumber. In a sinusoidally varying sound field the particle velocity is, from eq. (2.14), (2.21) Complex notation is used throughout this note except in the beginning of chapters 3 and 5 and at the end of chapter 6. Moreover, the equilibrium density of the medium is, for simplicity, written as D rather than D 0. 8

Example 2.1 Integration of eq. (2.10) over a small volume V defined by the closed surface S gives, with Gauss s theorem, which leads to the conclusion that the acoustic impedance (the complex ratio of sound pressure to volume velocity) of a cavity that is so small that we have the same sound pressure at all positions is (cf. eq. (2.9)), where is the volume velocity. (The minus sign is due to the fact that a positive divergence indicates a net flow out of the volume.) The air in the cavity behaves as a spring with a stiffness that is proportional to the static pressure and inversely proportional to the volume of the cavity. 3. PLANE WAVES IN DUCTS WITH RIGID WALLS The sound field in a uniform duct with rigid walls, sometimes referred to as an acoustic transmission line, is fairly simple in form at low frequencies where the cross-sectional dimensions are much smaller than the wavelength. As will be shown later, below a certain frequency that depends on the cross-sectional shape and dimensions of the duct only plane waves can propagate, from which it follows that the sound field is essentially onedimensional. This is the case in many situations of practical interest. In the following it is assumed that the sound field is one-dimensional. The wave equation (2.16) now assumes the form where z is the axial coordinate. It is easy to show that the general solution to this equation is 9 (3.1) (3.2) where f and g are continuous functions determined by the initial conditions. Since the argument of f is constant if (3.3) it can be seen that the first term of the general solution represents an acoustic disturbance of shape f(z) travelling undistorted and unattenuated in the z-direction with the speed c. The second term is a similar wave that travels in the opposite direction. Being independent of x and y both waves are plane waves. Example 3.1 If the tube is terminated by a rigid plane at z = 0 then the solution must satisfy the boundary condition u z (0, t) = 0 for all values of t. This is identical with requiring that the pressure gradient in the z-direction is zero at z = 0 for all values of t (cf. eq. (2.14)), and this is seen to be satisfied by the function

since It is apparently that there is perfect reflection at the rigid termination. See figure 3.1. Figure 3.1 Reflection at a rigid termination. (Adapted from ref. [4].) For several reasons it is advantageous to study the problem frequency by frequency, as mentioned in chapter 2. A one-dimensional harmonic sound field is governed by the onedimensional version of the Helmholtz equation (2.19), The general solution to this equation is (3.4) (3.5) where p + and p - are complex quantities. Equation (3.5) implies that the physical sound pressure is (3.6) which is seen to be in agreement with eq. (3.2), as of course it should be. The two terms of eq. (3.5) represent interfering plane waves with amplitudes and travelling in the z- direction and in the negative z-direction, respectively. Note that the sound field is spatially periodic with the period, (3.7) 10

where n is an integer, and 8, the wavelength, is the distance each of the two waves travels in one period of the sinusoidal signal, T = 1/f : (3.8) The particle velocity component in the z-direction follows from eqs. (2.21) and (3.5): (3.9) (The components in the x- and y-directions are zero, of course.) Note that (3.10) in a plane progressive wave travelling in the z-direction (corresponding to p - = 0), whereas (3.11) in a similar wave travelling in the opposite direction (corresponding to p + = 0). The quantity Dc is the characteristic impedance of the medium. 4 In the general case we have an interference field with a more complicated relation between the pressure and the particle velocity. The volume velocity, that is, the cross-sectional integral of the particle velocity, (3.12) is, as we shall see later, an important quantity in low-frequency modelling of ducts, just as it is an important quantity when dealing with sound radiation at low frequencies [3]. 3.1 The sound field in a duct terminated by an arbitrary acoustic impedance Figure 3.2 A tube terminated by a surface with the acoustic impedance Z a,t. Consider a duct terminated at z = 0 by a surface with a certain acoustic impedance, Z a,t, as sketched in figure 3.2. If a sound field is generated by a source, say, a piston, at a position z < 0, then the first term of eq. (3.5) can be interpreted as the incident part of the sound field, and the second term is the reflected part. It is obvious that. The complex ratio of the amplitudes at z = 0 is the reflection factor, (3.13) 4 The characteristic impedance of air under normal ambient conditions (101.3 kpa and 20/C) is approximately 413 kgm -2 s -1. 11

which completely characterises the acoustic properties of the surface for normal sound incidence. Equation (3.5) can now be written (3.14) It is easy to see from eq. (3.14) that the amplitude of the sound pressure varies with the position. The maximum value, (3.15) occurs at positions where the incident wave is in phase with the reflected wave (constructive interference), and the minimum value, (3.16) occurs at positions where the two waves are in antiphase (destructive interference). The ratio of the two quantities is the standing wave ratio, (3.17) From eqs. (3.9) and (3.14) it follows that the wave impedance (the complex ratio of the sound pressure to the particle velocity) 5 is (3.18) In figure 3.3 are shown the modulus and phase of the wave impedance as functions of the position for two different values of the standing wave ratio s. The boundary condition at z = 0 implies that It now follows that and (3.19) (3.20) (3.21) The latter equation shows that the quantity is important; this quantity may be called the characteristic impedance of the tube. Apparently, there is total reflection in phase (R 1) when the terminating surface is almost rigid ( ), there is total reflection in antiphase (R -1) when the terminating impedance is very small ( ), and there is no reflection (R = 0) when the terminating impedance equals the characteristic impedance of the tube, Dc/S. 5 This quantity is sometimes referred to as the specific impedance. 12

Figure 3.3 (a) Modulus (in Dc-units) and (b) phase of the wave impedance in a one-dimensional sound field for two different values of the standing wave ratio: ))), 24 db; - - -, 12 db. The sound intensity is (3.22) Equation (3.22) shows that the sound intensity is independent of the position, as of course it must be in a tube with rigid walls. Figure 3.4 Standing wave pattern in a tube terminated by a sample of a material with an absorption coefficient of 90% ( ))) ), 60% ( ) ) ) and 30% ( A A A ). Example 3.2 By definition, the absorption coefficient of the material that terminates the tube is the ratio of absorbed to incident sound power. Evidently, this quantity must take values between zero and unity. The incident sound power is the sound power associated with the incident wave, 13

and since the sound power absorbed by the material is it follows that the absorption coefficient is The modulus of the reflection factor can be expressed in terms of the standing wave ratio, cf. eq. (3.17), from which it can be seen that The standing wave ratio, and thus the absorption coefficient of the material at the end of the tube, can be measured by drawing a pressure microphone through the tube and recording the sound pressure as a function of the axial position. This is the classical method of determining normal incidence absorption coefficients. Example 3.3 It is possible to determine the reflection factor R and thus the absorption coefficient of the material at the end of the tube by measuring the transfer function between the signals from two pressure microphones mounted in the wall of the tube. Let where )z is the separation distance (see figure 3.5). The transfer function between these signals is Rearranging gives the expression which shows that R can be calculated from H 12, k, z and )z unless the distance between the microphones is a multiple of half a wavelength. Since this method does not require a traversing microphone, it is generally much faster than the classical method based on measuring the standing wave ratio. In practice the tube is driven with wide-band random noise, and the transfer function between the two pressure signals is measured using a dualchannel FFT analyser. By contrast the classical method requires driving the tube with one frequency at a time. Figure 3.5 The two-microphone technique of measuring absorption coefficients in a tube. 14

tube, Equation (3.18) makes it possible to determine the acoustic input impedance of the (3.23) where l is the length of the tube. Combining eqs. (3.21) and (3.23) gives the acoustic input impedance expressed in terms of the acoustic impedance of the material at the end of the tube: (3.24) It is apparent that the input impedance varies with the frequency unless Z a,t = Dc/S. Note also that Z a,i = Z a,t when kl is a multiple of B corresponding to the length of the tube being a multiple of half a wavelength. It is also interesting to observe that (3.25) when kl is an odd-numbered multiple of B/2 corresponding to the length of the tube being an odd-numbered multiple of a quarter of a wavelength; a tube of this length transforms a large impedance into a small one and vice versa. Example 3.4 The reflection factor equals unity if the tube is terminated by a rigid wall. The resulting interference pattern is called a standing wave. From eqs. (3.9) and (3.14), Note that the sound pressure and the particle velocity are in quadrature (90/ out of phase). Accordingly, the time-averaged sound intensity is zero at all positions in the tube. The input impedance is purely imaginary, and varies strongly with the frequency. At very low frequencies, where the tube is much shorter than the wavelength, the input impedance approaches the impedance of a small cavity, where V = Sl is the volume of the tube. Apparently, in the limit of low frequencies the impedance depends only on the volume. Compare with example 2.1; V/Dc 2 is the acoustic compliance of the cavity. Figure 3.6 shows the measured acoustic input impedance of a tube terminated by a rigid cap. The sound pressure and the particle velocity have been measured using a sound intensity probe. The measurement is not reliable below 100 Hz. 15

Figure 3.6 Modulus (in Dc-units) of the acoustic input impedance of a 50-cm long tube terminated by a rigid cap, measured using a sound intensity probe. When the terminating impedance is large, the acoustic input impedance of the tube becomes very small when l equals an odd-numbered multiple of a quarter of a wavelength. The corresponding discrete frequencies are the resonance frequencies of the system; the resonance frequencies are odd-numbered harmonics of the fundamental. Conversely, the input impedance becomes very large when l equals a multiple of half a wavelength, which occurs at the antiresonance frequencies of the system. One way of generating sound in the tube is by placing a vibrating piston at one end. Let the source be a surface at z = -l vibrating with the velocity. In the frequency range we are concerned with, the decisive property of the source is its volume velocity (this is expanded in section 6.3). The boundary condition at z = -l implies that (3.26) (3.27) Example 3.5 A semi-infinite tube is driven at the end by a piston with the volume velocity plane progressive wave with the sound pressure amplitude. The result is a Example 3.6 A tube of the length l terminated by a rigid surface at z = 0 is driven at z = - l by a piston with the vibrational velocity. Evidently and since, from example 3.4, it follows that from which it can be seen that 16

Note that the sound pressure goes towards infinity at the antiresonance frequencies owing to the fact that losses have been ignored. In practice there will always be some losses, and the source will have a finite acoustic impedance, which means that it cannot maintain its volume velocity if the sound pressure on its surface goes towards infinity. If the tube is driven with a small piston that is mounted in the wall, say at z = z 0, the boundary conditions at z = z 0 are (3.28) and (3.29) where and are the limits z 0 ± " with " 6 0, and is the volume velocity of the piston. The discontinuity in the volume velocity is due to the fact that the velocity is a vector. Example 3.7 A piston with the volume velocity is mounted at z = 0 in the wall of infinite tube. This source will generate plane waves travelling to the left and to the right: and From eqs. (3.28) and (3.29) it follows that share the volume velocity injected at z = 0.. In effect the right- and left-going waves Figure 3.7 A semi-infinite tube driven by a loudspeaker at an arbitrary position. Example 3.8 A semi-infinite tube is driven by a small loudspeaker with a volume velocity of mounted in the wall the distance l from a rigid termination (see figure 3.7). At positions between the source and the rigid surface (0 # z # l), and and since q(0) = 0 it can be seen that. To the right of the source (at z $ l), and 17

From eq. (3.28), and from eq. (3.29), from which it follows that the sound pressure between the source and the rigid surface is (a standing wave), and (a propagating wave) to the right of the source. Note that for z $ l at frequencies where l equals an odd-numbered multiple of a quarter of a wavelength. The physical explanation is that the sound field to the right of the source is the sum of the direct wave from the source and the left-going wave, reflected at z = 0, When l is an odd-numbered multiple of a quarter of a wavelength, the reflected wave has travelled an odd-numbered multiple of half a wavelength longer than the direct wave and is therefore in antiphase; therefore it cancels the direct wave. 3.2 Radiation of sound from an open-ended tube An open-ended tube is in effect terminated by the radiation impedance, which is the impedance that would load a piston at the end of the tube. The radiation impedance of a flanged tube (a tube with the open end in a large, rigid plane) is identical with the radiation impedance of a piston set in a large, plane, rigid baffle; the radiation impedance of an unflanged tube is similar to the radiation impedance of a small monopole in free space. As we shall see, at very low frequencies these quantities are much smaller than the characteristic impedance of the tube, which, with eq. (3.21), leads to the conclusion that R -1; there is a strong reflection in antiphase. 6 Accordingly, the input impedance at the other end of the tube exhibits minima when the (effective) length of the tube is a multiple of half a wavelength. The radiation impedance of a circular piston of the radius a mounted in an infinite baffle is [3] (3.30) where J 1 is the Bessel function and H 1 is the Struve function of first order. At low frequencies this expression can be approximated as follows [1], (3.31) The approximation is fair if ka < 0.5. At very low frequencies a considerable part of the incident sound power is reflected 6 A surface with the boundary condition is termed a pressure release surface. Such a surface gives total reflection in antiphase. 18

from the open end of the tube, as mentioned above. The transmission coefficient describes the small fraction of the incident sound power radiated into the open. From eq. (3.21) we have (3.32) from which it follows that the transmission coefficient J is given by (3.33) Note that the transmission coefficient to this approximation is independent of the imaginary part of the radiation impedance, even though the imaginary part is much larger than the real part. The imaginary part of the radiation impedance corresponds to the impedance of a mass of a volume of the medium of the size S)l, where (3.34) This becomes apparent if the radiation impedance is written as a mechanical impedance (the ratio of force to velocity): (3.35) The main effect of the imaginary part of the radiation impedance is to shift the position where reflection in antiphase occurs to a virtual plane outside the tube. This can be seen as follows. If we ignore the small real part of the radiation impedance eq. (3.32) becomes (3.36) corresponding to the input impedance of the tube being (3.37) and this is exactly the result we would obtain if a tube of the length l + )l was terminated by a surface with an acoustic impedance of zero (cf. eq. (3.24)). Equation (3.37) explains why the quantity )l is termed the end correction: the effective length of the tube is l + )l. Ignoring the real part of the radiation impedance is too coarse an approximation at the resonance frequencies, though. At these frequencies 19 (3.38)

which is small but not zero as implied by eq. (3.37). In a similar manner it can be shown that (3.39) at the antiresonance frequencies, and this impedance is large but not infinite as implied by eq. (3.37). In fact, the tube behaves as if it had a length of l + )l and were terminated by a small, purely real impedance, Re{Z a,r }. There is no simple way of calculating the radiation impedance of a thin-walled tube that radiates into open space without a flange. However, it can be shown [5] that in the lowfrequency limit, (3.40) corresponding to a transmission coefficient of and an end correction that takes a value of (3.41) (3.42) Apparently, a large flange tends to double the real part of the radiation impedance ) in agreement with the fact that the solid angle is (almost) halved by the flange, which implies a better impedance match. Note also that the real part of the radiation impedance of the unflanged tube, (3.43) is identical with the radiation resistance of a small pulsating sphere (see, e.g., ref. [3]). Figure 3.8 shows the input impedance of an open-ended tube. As the frequency is increased the impedance peaks and troughs become less and less pronounced. Figure 3.8 The input impedance of an open-ended tube. 20

Example 3.9 A source of high acoustic impedance, that is, a source that tends to maintain its volume velocity irrespective of the acoustic load, tends to excite the antiresonance frequencies of the tube. The generator in reed instruments, e.g. clarinets and oboes, and brass instruments, e.g. trumpets, may be regarded as a source of high impedance; therefore such instruments play at impedance peaks (antiresonance frequencies), which are odd-numbered harmonics of the fundamental. A tone produced by a clarinet therefore has a strong content of odd-numbered harmonics [4]. Example 3.10 A source of low acoustic impedance, that is, a source that tends to produce a given sound pressure at its surface irrespective of the acoustic load, tends to excite the resonance frequencies of the tube. The generator in flutes and organ pipes may be regarded as a source of low impedance; therefore flutes and open organ pipes play at impedance minima (resonance frequencies), which occur at all harmonics. It follows that a tone produced by a flute has even as well as odd harmonics [4]. Example 3.11 A tone produced by a closed organ pipe has only odd-numbered harmonics, because the resonances of a closed tube occur when the length equals an odd-numbered multiple of a quarter of a wavelength. Figure 3.9 Radiation from the open end of a tube. The sound field generated by an open-ended tube driven by a piston at the other end can be calculated as follows. From eq. (3.12) we can write where (3.44) (3.45) is the reflection factor at the virtual reflection plane z = 0 (cf. eq. (3.32)). The boundary condition at the piston implies that (3.46) where is the effective length of the tube. From eqs. (3.44) and (3.46), (3.47) (3.48) 21

which shows that the volume velocity is large when the cosine equals zero. In the frequency range we are concerned with the open end of the tube radiates as a monopole with the volume velocity, that is, the sound pressure at a distance r from the virtual opening is [3] (3.49) Example 3.12 The principle of reciprocity implies that the sound pressure at a given point r due to a monopole at another point r 0 is equal to the sound pressure generated by the monopole if the positions r and r 0 are interchanged. This leads to the conclusion that a monopole with the volume velocity radiating at the distance r from the open end of a tube would generate the sound pressure given by eq. (3.49) if the terminating surface at the other end of the tube is rigid. This result can also be obtained without employing the principle of reciprocity. The sound pressure at the distance r from the monopole is in the undisturbed medium. From Thévenin s theorem it now follows that the sound pressure at the open end of the tube is where is the input impedance of the closed tube (cf. example 3.4). There is a simple relation between the sound pressure at the rigid termination,, and the sound pressure at the open end, (cf. example 3.4). When these equations are combined eq. (3.49) results. 4. SOUND TRANSMISSION THROUGH COUPLED PIPES A silencer (or muffler) is a device designed to reduce the emission of noise from a pipe, for example the noise from the exhaust of an engine. Reactive silencers are composed of coupled tubes without absorbing material. They work by reflecting the sound from the source. Dissipative silencers absorb acoustic energy. There is no simple way of dealing with the sound field in a system of coupled tubes as the one sketched in figure 4.1 in the general case. However, at low frequencies where the sound field in each component part can be assumed to be one-dimensional, the problem is a fairly simple one. At each junction it can be assumed that the output sound pressure and volume velocity of one system equal the input sound pressure and volume velocity of the adjoining system; therefore the acoustic input impedance of the exit system is the acoustic load of the preceding system. 22

Figure 4.1 A system of coupled pipes. Example 4.1 Two tubes with cross-sectional areas S 1 and S 2 are connected, as sketched in figure 4.2. If the second tube is infinite, its input impedance is, from which it follows that the reflection factor at the junction seen from the first tube is (cf. eq. (3.21)). Note that R 1 if S 2 << S 1, in accordance with the fact that a sudden, strong area contraction almost acts as a rigid wall. Conversely, R - 1 if S 2 >> S 1 ; in this case the junction acts almost as a pressure-release surface. The transmission coefficient is It can be seen that the fraction of incident sound power transmitted from one tube to another is the same in both directions. It is physically obvious that it is unrealistic to expect the sound field to be strictly onedimensional in the immediate vicinity of a discontinuity as the one shown in figure 4.2. Nevertheless, if one s main concern is the transmission properties of the composite system it is usually a good approximation to assume that the sound field is one-dimensional (below a certain frequency determined by the cross-sectional dimensions), because the local disturbances correspond to higher-order modes that die out exponentially with the distance and have little effect on the overall transmission. Higher-order modes are analysed in chapter 6. Figure 4.2 Transmission and reflection at an area discontinuity. 4.1 The transmission matrix The low-frequency approximation implies that each subsystem is an acoustic two-port (or four-pole system) with two (and only two) unknown parameters, the complex amplitudes of two interfering waves travelling in opposite directions. Such a system can be described by its transmission (or four-pole) matrix, as follows, (4.1) 23