+ All Categories
Home > Documents > Intro to Acoustics

Intro to Acoustics

Date post: 06-Apr-2018
Category:
Upload: allen
View: 216 times
Download: 0 times
Share this document with a friend

of 49

Transcript
  • 8/3/2019 Intro to Acoustics

    1/49

    AAS

    1. INTRODUCTION

    1.1 Sources of sound

    These notes are concerned with the mechanisms of sound production and propagation.

    Music, calm speech, gently flowing streams or leaves fluttering in a breeze on a summer day

    are pleasant and desirable sounds. Noise, howling winds, explosions and angry voices are

    possibly less so. In all such cases we seek to understand the basic sources of the sound by

    careful analysis of the equations of motions. Most mechanical sources are very complex, very

    often involving ill-defined turbulent and perhaps combusting flows and their interactions

    with vibrating structures, and the energy released as sound is usually a tiny fraction of

    the structural and hydrodynamic energy of the source region. It is therefore important to

    develop robust mathematical models that reliably incorporate this general inefficiency of

    the sound generation mechanisms; small errors in source modelling can evidently lead to

    very large errors in acoustic prediction.

    To do this we consider a fluid that can be regarded as continuous and locally

    homogeneous at all levels of subdivision. At any time t and position x = (x1, x2, x3) the

    state of the fluid is defined when the velocity v and any two thermodynamic variables are

    specified for each of the fluid particles of which the fluid may be supposed to consist. The

    distinctive fluid property possessed by both liquids and gases is that these fluid particles

    can move freely relative to one another under the influence of applied forces or other

    externally imposed changes at the boundaries of the fluid. Five scalar partial differential

    equations are required to determine these motions. They are statements of conservation of

    mass, momentum and energy, and they are to be solved subject to appropriate boundaryand initial conditions, dependent on the problem at hand. These equations will be used to

    formulate and analyse a wide range of acoustic problems; our main task will be to simplify

    these problems to obtain a thorough understanding of source mechanisms together with a

    quantitative estimate of the radiated sound.

    M. S. Howe 1 1.1 Sources of sound

  • 8/3/2019 Intro to Acoustics

    2/49

    AAS

    1.2 Equations of motion of a fluid

    The state of a fluid at time t and position x = (x1, x2, x3) is defined when the velocity

    v and any two thermodynamic variables are specified. Five scalar equations are therefore

    required to determine the motion. These equations are statements of the conservation of

    mass, momentum and energy. They must normally be supplemented by a thermodynamic

    equation of state.

    1.2.1 The material derivative

    Let vi denote the component of fluid velocity v in the xi-direction, and consider the rate

    at which any function F(x, t) varies following the motion of a fluid particle. Let the particle

    be at x at time t, and at x + x a short time later at time t + t, where x = v(x, t)t + .At the new position of the fluid particle

    F(x + x, t + t) = F(x, t) + vjtF

    xj(x, t) + t

    F

    t(x, t) + ,

    where the repeated suffix j implies summation over j = 1, 2, 3. The limiting value of

    (F(x + x, t + t) F(x, t))/t as t 0 defines the material (or Lagrangian) derivative

    DF/Dt of F:

    DF

    Dt =

    F

    t + vj

    F

    xj

    F

    t + v F. (1.2.1)

    DF/Dt measures the time rate of change of F as seen by an observer moving with the fluid

    particle that occupies position x at time t.

    1.2.2 Equation of continuity

    A fluid particle of volume V and mass density has a total mass of V (V)(x, t),

    where x denotes the position of the centroid of V at time t (Figure 1.2.1). Conservation of

    mass requires that D(V)/Dt = 0, i.e. that

    1

    D

    Dt+

    1

    V

    DV

    Dt= 0. (1.2.2)

    Now 1VDVDt =

    1V

    Sv dS, where the integration is over the closed material surface S forming

    the boundary of V, on which the vector surface element dS is directed out of V. It is the

    M. S. Howe 2 1.2 Equations of motion

  • 8/3/2019 Intro to Acoustics

    3/49

    AAS

    fractional rate of increase of the volume of the fluid particle, and becomes equal to div v as

    V 0. In this limit (1.2.2) can therefore be cast in any of the following equivalent forms of

    the equation of continuity

    1 DDt + div v = 0,

    t

    + div(v) = 0,

    t +

    xj

    (vj) = 0.

    (1.2.3)

    Figure 1.2.1

    1.2.3 Momentum equation

    The momentum equation is derived in a similar manner by consideration of the rate of

    change of momentum D(Vv)/Dt (V)Dv/Dt of a fluid particle subject to pressure p

    and viscous forces acting on the surface S of V, and body forces F per unit volume within

    V. It is sufficient for our purposes to quote the form of the resulting Navier-Stokes equation

    Dv

    Dt= p curl +

    +

    4

    3

    div v + F, (1.2.4)

    where = curl v is the vorticity, and and are respectively the shear and bulk

    coefficients of viscosity, which generally vary with pressure, temperature and position in

    the fluid. Viscous forces are important predominantly close to solid boundaries, where the

    frictional drag is governed by the shear viscosity . It is then a good approximation toadopt the Stokesian model in which the contribution of the bulk viscosity is ignored.

    Values of , and = / (the kinematic viscosity) for air and water at 10C and one

    atmosphere pressure are given in the Table 1.2.1:

    M. S. Howe 3 1.2 Equations of motion

  • 8/3/2019 Intro to Acoustics

    4/49

    AAS

    kg/m3 kg/ms m2/s

    Air 1.23 1.764 105 1.433 105

    Water 1000 1.284 103 1.284 106

    Table 1.2.1

    1.2.4 Energy equation

    Consideration of the transfer of heat by molecular diffusion across the moving material

    boundary S of the fluid particle of Figure 1.2.1, and of the production of heat by frictional

    dissipation within its interior volume V leads to the energy equation

    TDsDt

    = 2

    eij 13

    ekkij2

    + e2kk + div

    T

    , (1.2.5)

    where T, s are respectively the temperature and specific entropy (i.e. entropy per unit

    mass), is the thermal conductivity of the fluid, and

    eij =1

    2

    vixj

    +vjxi

    (1.2.6)

    is the rate of strain tensor, which accounts for changes in shape of the fluid particle. Because

    ekk div v, the term involving the bulk coefficient of viscosity in (1.2.5) determines the

    dissipative production of heat during compressions and rarefactions of the fluid.

    To understand the significance of the rate of strain tensor, observe that the velocity v

    relative to the centroid of the moving fluid particle at vector distance x from its centroid,

    is given to first order by (see A.7)

    vi = xj

    vi

    xj xj

    1

    2

    vi

    xj+

    vj

    xi + xj

    1

    2

    vi

    xj

    vj

    xi

    =1

    2

    xi

    ejkx

    jx

    k

    +

    1

    2( x)i

    where eij and the vorticity are evaluated at the centroid. The term in therefore repre-

    sents relative rigid body rotation of the particle about the centroid, at angular velocity 12,

    M. S. Howe 4 1.2 Equations of motion

  • 8/3/2019 Intro to Acoustics

    5/49

    AAS

    with no change of shape. The gradient term, however, represents an irrotational distortion

    ofV, and is responsible for frictional forces and the conversion of mechanical energy into heat.

    1.2.5 Equation of state

    In the presence of velocity and pressure gradients a fluid cannot be in strict

    thermodynamic equilibrium, and thermodynamic variables require special interpretation.

    The density and the total internal energy e per unit mass can be defined in the usual

    way for a very small fluid particle without the need for thermodynamic equilibrium, such

    that and e are the mass and internal energy per unit volume. The pressure and all

    other thermodynamic quantities are then defined by means of the same functions of and e that would be used for a system in thermal equilibrium, and the relations between

    the thermodynamic variables are then the same as for a fluid in local thermodynamic

    equilibrium, defined by equations of the form

    p = p(, s), p = p(, T), s = s(, T), etc. (1.2.7)

    The equations of state (1.2.7) permit any thermodynamic variable to be expressed in terms

    of any two variables, such as the density and temperature, although in applications it may

    be more convenient to use other such equations. For sound propagation in an ideal fluid it

    is usual to neglect dissipation and to assume homentropic flow: s = so = constant. This

    permits the fluid motion to be determined from the equations of continuity and momentum

    and the equation of state p = p(, so), the energy equation being ignored. In more general

    situations it is necessary to retain the energy equation to account for coupling between

    macroscopic motions and the internal energy of the fluid.

    Note, however, that the thermodynamic pressure p = p(, e) defined in this way is

    generally no longer the sole source of normal stress on any surface drawn in the fluid.

    This is the case in a fluid of non-zero bulk viscosity ( = 0), whose molecules possess

    rotational (or other internal) degrees of freedom whose relaxation time required for the

    M. S. Howe 5 1.2 Equations of motion

  • 8/3/2019 Intro to Acoustics

    6/49

    AAS

    re-establishment of thermal equilibrium after, say, a compression, is large relative to the

    equilibration time of the translational degrees of freedom. When compressed (during an

    interval in which div v < 0) the temperature must rise, but the corresponding increase

    in the rotational energy lags slightly behind that of the molecular translational energy

    responsible for normal stress; the thermodynamic pressure p accordingly is smaller than the

    true normal stress (which equals p div v).

    1.2.6 Croccos equation

    The momentum equation (1.2.4) can be recast by introducing the specific enthalpy

    w = e + p/ of the fluid, in terms of which the first law of thermodynamics supplies therelation

    dw =dp

    + Tds. (1.2.8)

    The vector identity (v )v = v + 12v

    2

    and (1.2.8) permit the momentum equation

    to be put in Croccos form

    v

    t+ B = v + T s curl +

    + 43

    div v +

    F

    , (1.2.9)

    where = / and

    B = w + 12

    v2 (1.2.10)

    is the total enthalpy. In a perfect gas w = cpT = p/( 1), where cp is the specific

    heat at constant pressure and = cp/cv, cv being the specific heat at constant volume.

    Croccos equation finds application in the acoustics of turbulent, heat conducting flows.

    M. S. Howe 6 1.2 Equations of motion

  • 8/3/2019 Intro to Acoustics

    7/49

    AAS

    1.3 Sound waves in an ideal fluid

    The intensity of an acoustic pressure p in air (relative to the mean atmospheric pressure)

    is usually measured in decibels by the quantity

    20 log10

    |p|

    pref

    ,

    where the reference pressure pref = 2 105 Pa. Thus, p = po 1 atmosphere (= 105 Pa)

    is equivalent to 194 dB. A very loud sound 120 dB corresponds to

    p

    po

    2 105

    105 10(

    120

    20) = 2 104 1.

    Similarly, for a deafening sound of 160 dB, p/po 0.02. This corresponds to a pressure ofabout 0.3 lbs/in2, and is loud enough for nonlinear effects to begin to be important.

    The passage of a sound wave in the form of a pressure fluctuation is accompanied by a

    back-and-forth motion of the fluid in the direction of propagation at the acoustic particle

    velocity v, say. It will be seen (1.3.2) that

    acoustic particle velocity acoustic pressure

    o speed of sound,

    where o is the mean air density. The speed of sound in air is about 340 m/sec. Thus,

    v 5 cm/sec at 120 dB ; at 160 dB v 5 m/sec.

    1.3.1 The wave equation for an ideal fluid

    In most applications the acoustic pressure is very small relative to po, and sound

    propagation is studied by linearizing the equations of motion. We consider first the

    simplest case of sound propagation in an ideal fluid i.e. a homogeneous, inviscid,

    non-heat-conducting fluid of mean pressure po and density o, which is at rest in the

    absence of the sound. The energy equation (1.2.5) implies that Ds/Dt = 0 so that sound

    propagation is homentropic (adiabatic) with s = so s(po, o) = constant throughout fluid.

    The implication is that, in an ideal fluid there is negligible dissipation of the organized

    M. S. Howe 7 1.3 Ideal fluid

  • 8/3/2019 Intro to Acoustics

    8/49

    AAS

    mechanical energy of the sound by heat and momentum transfer by molecular diffusion

    between neighboring fluid particles.

    The departures of the pressure and density from their undisturbed values are denoted

    by p, where p/po 1, /o < 1. The linearized form of the momentum equation

    (1.2.4) for an ideal fluid ( = = 0) then becomes

    ov

    t+ p = F. (1.3.1)

    Before linearizing the continuity equation (1.2.3) it is useful to make an artificial

    generalization by inserting a volume source distribution q(x, t) on the right hand side:

    1

    D

    Dt+ div v = q; (1.3.2)

    q is the rate of increase of fluid volume per unit volume of the fluid, and might represent, for

    example, the effect of volume pulsations of a small body in the fluid ( 1.4). The linearized

    equation is then

    1

    o

    t+ div v = q. (1.3.3)

    Eliminate v between (1.3.1) and (1.3.3):

    2

    t2 2p = o

    q

    t div F. (1.3.4)

    An equation determining the pressure p alone in terms ofq and F is obtained by invoking

    the homentropic approximation p = p(, so), where po = p(o, so) in the undisturbed state.

    Therefore

    po + p = p(o +

    , so) p(o, so) + p

    (, so), (1.3.5)

    where the derivative is evaluated at the undisturbed value o of the density. It has the

    dimensions of (velocity)2 and its square root defines the speed of sound

    co =

    p

    s

    , (1.3.6)

    M. S. Howe 8 1.3 Ideal fluid

  • 8/3/2019 Intro to Acoustics

    9/49

    AAS

    where differentiation is performed at s = so, and evaluated at = o. In air co 340m/s;

    in water co 1500 m/s.

    From (1.3.5): = p/c2o, and substitution for in (1.3.4) yields the inhomogeneous

    wave equation 1

    c2o

    2

    t2 2

    p = o

    q

    t div F, (1.3.7)

    where the prime on the acoustic pressure has been discarded. This equation governs the

    production of sound waves by the volume source q and the force F. When these terms are

    absent the equation describes sound propagation from sources on the boundaries of the

    fluid, such as the vibrating cone of a loudspeaker.

    The volume source q and (with the exception of gravity) the body force F would never

    appear in a complete description of sound generation in a real fluid. They are introduced

    only when we think we understand how to model mathematically the real sources of

    sound in terms of idealized volume sources and forces. In general this can be a dangerous

    procedure because, as we shall see, small errors in specifying the sources of sound in a fluid

    often result in very large errors in the predicted sound. This is because only a tiny fraction

    of the available energy of a vibrating fluid or structure actually radiates away as sound.

    When F = 0 equation (1.3.1) implies the existence of a velocity potential such that

    v = , in terms of which the perturbation pressure is given by

    p = o

    t. (1.3.8)

    It follows from this and (1.3.7) (with F = 0) that the velocity potential is the solution of

    1

    c2

    o

    2

    t2

    2 = q(x, t). (1.3.9)Causality can be invoked to justify the neglect of any a time-independent constants of

    integration. Equation (1.3.9) is the wave equation of classical acoustics.

    In the propagation zone, where the source terms q = 0, F = 0, the velocity v and the

    perturbations in p, (and in other thermodynamic quantities such as the temperature T,

    M. S. Howe 9 1.3 Ideal fluid

  • 8/3/2019 Intro to Acoustics

    10/49

    AAS

    internal energy e and enthalpy w, but not the specific entropy s, which remains constant and

    equal to so) propagate as sound governed by the homogeneous form of (1.3.9). The velocity

    fluctuation v produced by the passage of the wave is the acoustic particle velocity.

    1.3.2 Plane waves

    A plane acoustic wave propagating in the x-direction satisfies

    1

    c2o

    2

    t2

    2

    x2

    = 0, (1.3.10)

    which has the general solution (DAlembert 1747)

    =

    t xco

    +

    t + x

    co

    , (1.3.11)

    where , are arbitrary functions that respectively represent waves travelling at speed

    co without change of form in the positive and negative x-directions. The acoustic particle

    velocity v = is parallel to the propagation direction (the waves are longitudinal).

    The solutions (1.3.11) and the linearized (source-free) equations of motion can be used

    to show that fluctuations in v, p, , T and w in a plane wave propagating parallel to the

    x-axis are related by

    v = p

    oco, =

    p

    c2o, T =

    p

    ocp, w =

    p

    o, (1.3.12)

    where p is the acoustic pressure, cp is the specific heat at constant pressure, and the sign

    is taken according as the wave propagates in the positive or negative x-direction.

    1.3.3 Speed of sound

    In a perfect gas p = RT and s = cv ln(p/), where R = cp cv is the gas constant,

    and in the linearized approximation

    co =

    po/o =

    RTo. (1.3.13)

    M. S. Howe 10 1.3 Ideal fluid

  • 8/3/2019 Intro to Acoustics

    11/49

    AAS

    Typical approximate speeds of sound in air and in water, and the corresponding acoustic

    wavelength = co/f at a frequency of f = 1 kHz are given in Table 1.3.1

    co at 1 kHzm/s f/s km/h mph metres feet

    Air 340 1100 1225 750 0.3 1

    Water 1500 5000 5400 3400 1.5 5

    Table 1.3.1 Speed of sound and acoustic wavelength

    Example 1. Waves in a uniform tube generated by an oscillating piston The endx = 0 of an infinitely long, uniform tube is closed by a smoothly sliding piston executing

    small amplitude normal oscillations at velocity uo(t) (Figure 1.3.1a). If x increases along

    the tube, linear acoustic theory and the radiation condition require that p = (t x/co).

    At x = 0 the velocities of the fluid and piston are the same, so that

    uo(t) (t)

    oco, . . p = ocouo(t x/co) for x > 0.

    In practice a solution of this kind, where energy is confined by the tube to propagate in

    waves of constant cross-section, becomes progressively invalid as x increases, because of the

    accumulation of small effects of flow nonlinearity. Nonlinear analysis reveals that at a point

    in the wave where the particle velocity is v the wave actually propagates at speed co + v,

    so that wave elements where v is large and positive produce wave steepening, resulting

    ultimately in the formation of shock waves. This type of behaviour is important, for

    example, for waves generated in a long railway tunnel by the piston effect of an entering

    high-speed train.

    Example 2. Reflection at a closed end (Figure 1.3.1b) Let the plane wave

    p = pI(t x/co) approach from x < 0 the closed, rigid end at x = 0 of a uniform,

    M. S. Howe 11 1.3 Ideal fluid

  • 8/3/2019 Intro to Acoustics

    12/49

    AAS

    semi-infinite tube. The reflected pressure pR(t + x/co) is determined by the condition that

    the (normal) fluid velocity must vanish at x = 0. Therefore pI(t)/oco pR(t)/oco = 0,

    and the overall pressure within the tube is given by

    p = pI(t x/co) + pI(t + x/co), x < 0.

    Reflection at the rigid end causes pressure doubling at the wall where p = 2pI(t).

    Figure 1.3.1

    Example 3. Reflection at an open end (Figure 1.3.1c) When the wavelength of the

    sound is large compared to the radius R of an open ended circular cylindrical tube, the first

    approximation to the condition satisfied by the acoustic pressure at the open end ( x = 0) is

    that the overall pressure p = 0. Indeed, because of the free expansion of the fluid outside

    the tube, the pressure outside may be assumed to vanish compared to the incident pressure

    M. S. Howe 12 1.3 Ideal fluid

  • 8/3/2019 Intro to Acoustics

    13/49

    AAS

    pI(t x/co) pI(t). Then, integration of the linearized momentum equation over the

    fluid contained in a spherical region V of radius Rs, where R Rs = the acoustic

    wavelength, reveals that

    R2p t

    V

    ov1 d3x

    f V pI(t)co

    ,

    where v1 O(pI(t)/oco) is the particle velocity parallel to the tube, p is the net pressure

    within the tube close to the open end, and f is the frequency of the sound.

    Therefore, near the open end

    p fV

    R2copI(t)

    R3sR2

    pI(t) pI(t) as =cof

    .

    Thus, relative to the incident pressure, the pressure at the open end when R may be

    assumed to vanish. The pressure wave reflected back into the tube at the end is therefore

    approximately pI(t + x/co),

    p pI(t x/co) pI(t + x/co), x < 0,

    and the acoustic particle velocity in the mouth of the tube 2pI(t)/oco, twice that

    attributable to the incident wave alone.

    Example 4. Low frequency resonant oscillations in a pipe with open ends (Figure

    1.3.1d) A pressure wave of complex amplitude p and radian frequency propagating in

    the x-direction has the representation p = Re{pei(tx/co)}. Therefore, the combination

    p = Repei(tx/co) pei(t+x/co)

    vanishes at x = 0, and vanishes also at x = for those frequencies satisfying

    ei/co ei/co = 0, i.e. sin(ko) = 0

    where ko = /co is called the acoustic wavenumber. This equation accordingly determines

    the resonance frequencies of an open-ended tube of length in the low frequency

    approximation in which the pressure is assumed to vanish at the ends, viz

    f =

    2

    n co2

    , n = 1, 2, 3, . . . .

    M. S. Howe 13 1.3 Ideal fluid

  • 8/3/2019 Intro to Acoustics

    14/49

    AAS

    Thus, the minimum resonance frequency of a pipe of length = 1m is about 170Hz

    (co 340 m/s in air), whose wavelength co/f = 2 = 2 m does not depend on the speed of

    sound.

    The simple, one dimensional theory neglects energy losses from the ends of the tube

    by radiation into the ambient atmosphere, and neglects also viscous and thermal losses

    in acoustic boundary layers at the walls of the tube. These cause the wave amplitude

    to decay after several wave periods, so that resonant oscillations within the tube actually

    persist only for a finite time after the source of excitation is removed.

    M. S. Howe 14 1.3 Ideal fluid

  • 8/3/2019 Intro to Acoustics

    15/49

    AAS

    1.4 Low frequency pulsations of a sphere

    Small amplitude, irrotational motion produced by radial pulsations of a sphere of mean

    radius a with centre at the origin satisfy the homogeneous equation

    1

    c2o

    2

    t2 2 = 0, r = |x| > a. (1.4.1)

    If the oscillations occur at frequency (proportional to eit, for example), then /t ,

    and very close to the sphere the two terms on the left of this equation are respectively of

    orders k2o and /a2, where ko = /co is called the acoustic wavenumber. Ifkoa 1 the

    sphere is said to be acoustically compact. The characteristic wavelength 2/ko of

    the sound produced by the pulsations is then much larger than the radius a. More generally,a body is said to be acoustically compact when its characteristic dimension is small

    compared to the wavelengths of the sound waves it is producing or with which it interacts.

    In this limit the unsteady motion in the immediate neighbourhood of the sphere is governed

    by the Laplace equation obtained by discarding the first term on the left of (1.4.1):

    2 = 0, r > a. (1.4.2)

    This is just the continuity equation div v = 0 for incompressible flow. The corresponding

    pressure and density perturbations p and satisfy

    p

    c2o.

    In an incompressible fluid the pressure changes by the action of external forces (moving

    boundaries, etc), but the density must remain fixed. Thus, the formal limit of incompressible

    flow corresponds to setting co

    =

    .

    M. S. Howe 15 1.4 Pulsating sphere

  • 8/3/2019 Intro to Acoustics

    16/49

    AAS

    1.4.1 Pulsating sphere in incompressible fluid

    Let the normal velocity on the mean position r = a of the surface of the sphere be vn(t).

    When the motion is incompressible we have to solve

    2 = 0, r > a,

    /r = vn(t), r = a

    where r = |x|.

    Figure 1.4.1

    The solution must be radially symmetric, so that

    2 =1

    r2

    r

    r2

    r

    = 0, r > a.

    Hence

    =A

    r+ B, where A A(t), B B(t) are functions of t.

    B(t) can be discarded, because the pressure fluctuations ( o/t) must vanish as

    r , and a constant value of B has no physical significance. Applying the condition

    /r = vn, r = a we then find

    = a2vn(t)

    r, r > a. (1.4.3)

    Thus, the pressure

    p = o

    t= o

    a2

    r

    dvndt

    (t)

    M. S. Howe 16 1.4 Pulsating sphere

  • 8/3/2019 Intro to Acoustics

    17/49

    AAS

    decays like 1/r with distance from the sphere, and exhibits the unphysical characteristic

    of changing instantaneously everywhere when dvn/dt changes its value. For any time t

    the volume flux q(t) of fluid is the same across any closed surface enclosing the sphere.

    Evaluating it for any sphere S of radius r > a we find

    q(t) =S

    dS = 4a2vn(t),

    and we may also write

    =q(t)

    4r, r > a. (1.4.4)

    1.4.2 Point source in incompressible fluid

    The incompressible motion generated by a volume point source of strength q(t)

    concentrated at the origin is the solution of equation (1.3.9) with co = and

    q(x, t) = q(t)(x):

    2 = q(t)(x), where (x) = (x1)(x2)(x3). (1.4.5)

    The solution must be radially symmetric and given by

    =A

    r, for r > 0. (1.4.6)

    To find A equation (1.4.5) is integrated over the interior of a sphere of radius r = R > 0,

    and the divergence theorem is applied on the left:r a = radius of the sphere. This

    indicates that when we are interested in modelling the effect of a pulsating sphere at large

    M. S. Howe 17 1.4 Pulsating sphere

  • 8/3/2019 Intro to Acoustics

    18/49

    AAS

    distances r a, it is permissible to replace the sphere by a point source (a monopole)

    of the same strength q(t) = rate of change of the volume of the sphere. This conclusion is

    valid for any pulsating body, not just a sphere. However, it is not necessarily a good model

    (especially when we come to examine the production of sound by a pulsating body) in the

    presence of a mean fluid flow past the sphere.

    The solution (1.4.6) for the point source is strictly valid only for r > 0, where it satisfies

    2 = 0. What happens as r 0, where its value is actually undefined? To answer this

    question we write the solution in the form

    = lim0

    q(t)

    4(r

    2

    +

    2

    )

    1

    2

    , > 0, in which case 2 = lim0

    32q(t)

    4(r

    2

    +

    2

    )

    5

    2

    .

    The last limit is just equal to q(t)(x). Indeed when is small 32/4(r2 + 2)5

    2 is also

    small except close to r = 0, where it attains a large maximum 3/43. Therefore, for any

    smoothly varying test function f(x) and any volume V enclosing the origin

    lim0

    V

    32f(x)d3x

    4(r2 + 2)5

    2

    = f(0)lim0

    32d3x

    4(r2 + 2)5

    2

    = f(0)0

    32r2dr

    (r2 + 2)5

    2

    = f(0),

    where the value of the last integral is independent of . This is the defining property of the

    three-dimensional -function.

    Thus the correct interpretation of the solution

    =1

    4rof 2 = (x) (1.4.7)

    for a unit point source (q = 1) is

    1

    4r= lim

    0

    1

    4(r2 + 2)1

    2

    , r 0, (1.4.8)

    where

    2

    1

    4r

    = lim

    02

    1

    4(r2 + 2)1

    2

    = lim

    0

    32

    4(r2 + 2)5

    2

    = (x). (1.4.9)

    M. S. Howe 18 1.4 Pulsating sphere

  • 8/3/2019 Intro to Acoustics

    19/49

    AAS

    1.4.3 Low frequency pulsations of a sphere in compressible fluid

    Let us next calculate the radially symmetric sound produced by the acoustically compact

    pulsating sphere of1.4.1. Setting r = |x| and observing that radial symmetry implies that

    2 1

    r2

    r

    r2

    r

    1

    r

    2

    r2(r) ,

    it follows that (1.4.1) reduces to the one dimensional wave equation for r

    1

    c2o

    2

    t2(r)

    2

    r2(r) = 0, when r > a. (1.4.10)

    The general solution r = (t r/co) + (t + r/co), for arbitrary functions and , yields

    =

    t r

    co

    r+

    t +

    r

    co

    r, r > a. (1.4.11)

    The terms on the right respectively represent spherically symmetric disturbances

    propagating in the directions of increasing and decreasing values of r at the speed of

    sound co. Causality requires the incoming wave to be omitted, i.e. that = 0, because

    (t + r/co)/r necessarily represents sound arriving from r = and must be absent for

    pulsations that started at some finite time in the past. This statement of the causality

    principle is equivalent to imposing a radiation condition that sound waves must radiate

    away from their source.

    The outgoing wave function is determined from the boundary condition /r = vn(t)

    at r = a, which gives

    vn(t) =

    r

    t rco

    r

    r=a

    = 1

    a2

    t a

    co

    1

    aco

    t

    t

    a

    co

    . (1.4.12)

    But for a typical component of (t) eit of radian frequency ,

    ei(ta/co) eit anda

    co

    t koa 1,

    when the sphere is compact (koa 1). Therefore, the time delay a/co on the right of

    (1.4.12) can be neglected and the time-derivative term discarded, giving (t) a2vn(t),

    M. S. Howe 19 1.4 Pulsating sphere

  • 8/3/2019 Intro to Acoustics

    20/49

    AAS

    and

    q(t r/co)

    4r, r > a (1.4.13)

    where q(t) = 4a2vn(t) is the rate of volume outflow from the sphere. The acoustic potential

    at time t at a distant point r is seen to be the same as for incompressible flow (equation

    (1.4.4)) except that it is delayed by the time of travel r/co of sound from the sphere.

    This result is typical of compact, pulsating bodies; the sound can always be expressed in

    form (1.4.13) terms of the pulsational volumetric flow rate q(t).

    M. S. Howe 20 1.4 Pulsating sphere

  • 8/3/2019 Intro to Acoustics

    21/49

    AAS

    1.5 Sound produced by an impulsive point source

    The impulsive point source of strength q = (x)(t) is non-zero for an infinitesimal

    time at t = 0. The usual convention in acoustics, however, is to reverse the sign of the

    source (so that q(x, t) = (x)(t)), and to consider the corresponding inhomogeneous

    wave equation (1.3.9) in the form

    1

    c2o

    2

    t2 2

    = (x)(t). (1.5.1)

    Because the source vanishes for t < 0 we are interested only in the causal solution, which is

    non-zero only for t > 0.

    The solution is radially symmetric and of outgoing wave form, so that we can put

    = (t r/co)

    rfor r = |x| > 0. (1.5.2)

    The functional form of can be determined by the method of 1.4.2, or more simply by

    noting that

    (t r/co)

    r= lim

    0

    (t)

    (r2 + 2)1

    2

    when r 0, and therefore that temporal derivatives /t become negligible comparedto spatial derivatives /r. In other words the solution must resemble that for an

    incompressible fluid very close to the source. Hence we must have (t) = (t)/4 and the

    causal solution of (1.5.1) becomes

    (x, t) =1

    4r

    t r

    co

    1

    4|x|

    t

    |x|

    co

    . (1.5.3)

    The sound wave consists of a singular spherical pulse that is non-zero only on the surface of

    the sphere r = cot > 0 expanding at the speed of sound; it vanishes everywhere for t < 0.

    M. S. Howe 21 1.5 Impulsive point source

  • 8/3/2019 Intro to Acoustics

    22/49

    AAS

    1.6 Free space Greens function

    The free space Greens function G(x, y, t , ) is the causal solution of the wave equation

    generated by the impulsive point source (x y)(t ), located at the point x = y at

    time t = . The formula for G is obtained from the solution (1.5.3) for a source at x = 0 at

    t = 0 simply by replacing x by x y and t by t . In other words, if

    1

    c2o

    2

    t2 2

    G = (x y)(t ), where G = 0 for t < , (1.6.1)

    then

    G(x, y, t , ) =1

    4|x y|

    t

    |x y|

    co

    . (1.6.2)

    This represents an impulsive, spherically symmetric wave expanding from the source at y

    at the speed of sound. The wave amplitude decreases inversely with distance |x y| from

    the source point y.

    1.6.1 The retarded potential

    Greens function is the fundamental building block for solutions of the inhomogeneous

    wave equation (1.3.7) of linear acoustics in an unbounded medium. Let us write this

    equation in the form 1

    c2o

    2

    t2 2

    p = F(x, t), (1.6.3)

    where the generalized source F(x, t) is assumed to be generating waves that propagate

    away from the source region, in accordance with the radiation condition.

    This source distribution can be regarded as a distribution of impulsive point sources of

    the type on the right of equation (1.6.1), because

    F(x, t) =

    F(y, )(x y)(t )d3yd.

    The outgoing wave solution for each constituent source of strength

    F(y, )(x y)(t )d3yd is F(y, )G(x, y, t , )d3yd,

    M. S. Howe 22 1.6 Greens function

  • 8/3/2019 Intro to Acoustics

    23/49

    AAS

    so that by adding these individual contributions we obtain

    p(x, t) =

    F(y, )G(x, y, t , )d3yd (1.6.4)

    =1

    4

    F(y, )

    |x y|

    t

    |x y|

    co

    d3yd (1.6.5)

    i.e. p(x, t) =1

    4

    F

    y, t |xy|co

    |x y|

    d3y. (1.6.6)

    The integral formula (1.6.6) is called a retarded potential; it represents the pressure at

    position x and time t as a linear superposition of contributions from sources at positions

    y which radiated at the earlier times t |x y|/co, |x y|/co being the time of travel of

    sound waves from y to x.

    1.6.2 Greens function in one or two space dimensions: method of descent

    The Greens function for plane waves that propagate in one dimension (in a uniform

    duct, for example) say parallel to the x1-axis, is the causal solution G(x1, y1, t , ) of

    1c2o

    2

    t2

    2

    x21

    G = (x1 y1)(t ), where G = 0 for t < . (1.6.7)

    This equation can be obtained formally by integrating the corresponding three dimensional

    equation (1.6.1) over the plane of uniformity < x2, x3 < . This is Hadamards

    (1952) method of descent to a lower space dimension. Using the formula (1.6.2) for G in

    three dimensions, we find that (1.6.7) is satisfied by

    G(x1, y1, t , ) =

    1

    4|x y|

    t

    |x y|

    co

    dy2dy3

    =0

    t

    |x1 y1|2 + 2

    co

    d

    2

    |x1 y1|2 + 2

    =co2

    H

    t

    |x1 y1|

    co

    , (1.6.8)

    M. S. Howe 23 1.6 Greens function

  • 8/3/2019 Intro to Acoustics

    24/49

    AAS

    where H(x) is the Heaviside unit step function (= 1, 0 according as x >< 0; see Appendix B).

    In two dimensions sound waves propagate cylindrically as functions of (x1, x2) and t,

    and Greens function can be found by descent, by integrating the three dimensional formula

    over < x3 < to obtain

    G(x, y, t , ) =H(t |x y|/co)

    2

    (t )2 |x y|2/c2o, x = (x1, x2), y = (y1, y2), (1.6.9)

    the causal solution of

    1

    c2o

    2

    t2 2

    G = (x1 y1)(x2 y2)(t ), G = 0 for t < .

    In three dimensions G consists of a spherically spreading singular pulse that vanishes

    everywhere except at the wavefront. The corresponding one dimensional Greens function

    is finite and consists of two simple discontinuities propagating in both directions from the

    source at the speed of sound, to the rear of which the amplitude is constant and equal to

    co/2. The behaviour of G in two dimensions exhibits certain intermediate characteristics:

    the wavefront consists of a circular cylindrical singular pulse radiating outwards from the

    source at the speed of sound, but followed by a slowly decaying tail extending back to the

    source point where its amplitude decreases like 1/(t ). From the view point of an observer

    in three dimensions, the two dimensional source (x1 y1)(x2 y2)(t ) is equivalent to

    a uniform, infinitely long line source (parallel to the x3 axis). The tail can be attributed

    to the arrival of sound from distant points on this line source, which persists for all time

    after the passage of the wave front, which can be regarded as produced by components of

    the line source in the immediate neighbourhood of its intersection with the x1x2 plane.

    M. S. Howe 24 1.6 Greens function

  • 8/3/2019 Intro to Acoustics

    25/49

    AAS

    1.7 Initial value problem for the wave equation

    Our first application of the retarded potential integral is to the solution of Cauchys

    problem, where it is required to calculate the sound at times t > 0 in terms of the state

    of the acoustic medium at t = 0. This includes, for example, the problem of determining

    the sound produced by the sudden rupture of a closed material envelope containing air at

    high pressure (a bursting balloon). Alternatively, the state of a system of sound waves

    generated by sources in the distant past might be specified and it is desired to determine

    their subsequent propagation.

    The nature of required initial data at t = 0, say, may be deduced directly from the

    homogeneous form of the wave equation (1.6.3), whose solution is required for t > 0. To do

    this the equation is multiplied by the Heaviside function H(t), making use of the identity

    H(t)2p

    t2=

    2

    t2

    pH(t)

    t

    (t)p

    (t)

    p

    t.

    Then equation (1.6.3), with F(x, t) 0, becomes

    1

    c2o

    2

    t2 2

    (pH) =

    1

    c2o

    t

    (t)p

    +

    1

    c2o(t)

    p

    t. (1.7.1)

    This equation is formally valid for all time < t < , with pH(t) 0 for t < 0.

    The outgoing wave solution calculated using the retarded potential (1.6.6) determines

    p(x, t)H(t) p(x, t) for t > 0 in terms of the initial pressure and velocity distributions:

    p = f(x),p

    t= oc

    2odiv v = g(x) at t = 0, (1.7.2)

    where f(x), g(x) represent the prescribed initial values (p/t = oc2odiv v is just the

    linearised continuity equation).

    To evaluate the retarded potential integral for t > 0 we start from (1.6.5):

    p(x, t) =1

    4c2o

    f(y)()

    + g(y)()

    t

    |x y|

    co

    d3y d

    |x y|. (1.7.3)

    M. S. Howe 25 1.7 Initial value problem

  • 8/3/2019 Intro to Acoustics

    26/49

    AAS

    Introduce polar coordinates (r = |x y|, , ), with origin at the observer position x, so

    that d3y = r2dr d, where d = sin dd is the solid angle element. Then the integral

    involving f(y) becomes

    1

    4c2o

    f(r, )()

    t r

    co

    rdrd =

    1

    4c2o

    t

    f(r, )

    t

    r

    co

    rdrd

    =

    t

    tf(cot, )

    d

    4

    =

    t

    tf(cot)

    ,

    where

    f(cot) =

    f(cot, )d

    4

    is the mean value of f(y) f(r, ) on the surface of a sphere of radius r = cot centred on

    the observer position x (P in Figure 1.7.1).

    Combining this with a similar calculation for the term in g(y), leads to Poissons (1819)

    solution

    p(x, t) =

    t

    tf(cot)

    + tg(cot). (1.7.4)

    Figure 1.7.1

    If the initial values f(x), g(x) of the pressure and its time derivative are non-zero only

    within a finite region bounded by a closed surface S (Figure 1.7.1), the mean values on the

    right of (1.7.4) vanish except when the expanding spherical surface r = cot cuts across S.

    The pressure perturbation at P outside S is therefore non-zero only for r1 < cot < r2, where

    M. S. Howe 26 1.7 Initial value problem

  • 8/3/2019 Intro to Acoustics

    27/49

    AAS

    r1, r2 are the respective radii of the smallest and largest spheres centred on P that just

    touch S. The acoustic pressure radiating from S is therefore non-zero only within a shell-like

    region of space; at time t the outer surface of this shell is the envelope of the family of

    spheres of radius cot whose centres lie on S. Equation (1.7.4) also gives the solution for an

    observer within S. For sufficiently small time the inner envelope of the expanding family

    of spheres forms a wavefront collapsing into the interior of S this eventually crosses itself,

    emerges from the other side of S and expands to form the inner boundary of the radiating

    shell.

    Note that the right hand side of (1.7.1) can be cast in the form of the general source of

    the linear acoustic equation (1.3.7)

    oq

    t div F

    where q(x, t), F(x, t) are impulsive volume source and body force distributions

    q(x, t) =1

    oc2op(x, 0)(t), F(x, t) = ov(x, 0)(t), (1.7.5)

    and where the formula for F is obtained from the linearised continuity equation

    1

    c2o

    p

    t + div(ov) = 0.

    Also, when the initial disturbance is non-zero only within a finite region bounded by the

    surface S, the linearised momentum equation (1.3.1) implies that at any point x outside

    S 0 p(x, t)dt = o[v(x, t)]

    t=0 0, because v = 0 before and after the passage of the

    sound wave. Therefore 0

    p(x, t)dt = 0 for all x, (1.7.6)

    because the integral must vanish at infinitely large distances from S. Thus, the acoustic

    pressure variation at x during the time interval r1/co < t < r2/co occupied by the wave

    must always involve equal and opposite net compressions and rarefactions. This conclusion

    is true for waves propagating in three and two dimensions, but not for one-dimensional

    propagation. It is illustrated by the following example.

    M. S. Howe 27 1.7 Initial value problem

  • 8/3/2019 Intro to Acoustics

    28/49

    AAS

    1.7.1 Sound radiated from a spherical region of initial high pressure

    Let the initial uniform high pressure f(x) = p0 > 0 be confined to stationary fluid in the

    interior of a sphere S of radius a, so that g(x) = 0. Evidently, for the exterior point P in

    Figure 1.7.2 at distance r from the centre O of the sphere, r1 = r a, r2 = r + a, and the

    pressure at P is non-zero only for r a < cot < r + a. During this time

    f(cot) =p04

    ,

    where =20 d

    (t)0 sin d = 2{1 cos (t)} is the solid angle subtended at P by the

    spherical cap AB formed by the intersection of S and the sphere of radius cot centred on P.

    Hence, using the cosine formula (a2

    = r2

    + c2ot2

    2rcot cos ) for the triangle OAP, we find

    f(cot) =p0

    4rcot

    a2 (cot r)

    2

    H(cot r + a) H(cot r a)

    ,

    and therefore

    p(r, t) =

    t

    tf(cot)

    =

    p02r

    (cot r)

    H(cot r + a) H(cot r a)

    , r > a, (1.7.7)

    which represents an outgoing spherical wave confined to the shell cot a < r < cot + a that

    decreases in amplitude like 1/r.

    Figure 1.7.2

    A similar calculation performed when P lies within the initial high pressure region

    (r < a) gives

    p(r, t) = p0

    1

    (cot + r)

    2rH(cot + r a) +

    (cot r)

    2rH(cot r a)

    , r < a. (1.7.8)

    M. S. Howe 28 1.7 Initial value problem

  • 8/3/2019 Intro to Acoustics

    29/49

    AAS

    The pressure at r inside S remains equal to p0 until the arrival of the rarefaction wave (the

    second term in the large brackets) at time t = (a r)/co, which is subsequently reflected

    from the centre r = 0 as a compression wave, after the passage of which the pressure is

    reduced to zero.

    The sequence of events is illustrated in Figure 1.7.3, where the nondimensional pressure

    p/p0 is plotted against r/a for a set of increasing values of cot/a. The perturbation

    pressure is initially uniform within S at cot/a = 0 and vanishes elsewhere. Compression

    and rarefaction waves radiate respectively into the exterior and interior of S as cot/a

    increases towards 1 . When cot/a 1 the negative rarefaction peak becomes very large, and

    ultimately p/p0 (cot/a 1) at r = 0. Long before this happens in a real fluid, however,the large rarefaction is suppressed by nonlinear actions that increase the propagation speed

    of the higher pressure sections of the inward propagating wave, thereby inhibiting the

    formation of a deep negative pressure. After reflection at r = 0 at cot/a = 1 the whole

    disturbance becomes outgoing of N-wave profile occupying a shell of thickness 2a, the

    compression wavefront being at r = cot + a and the pressure vanishing within the shell

    at r = cot. Spherical spreading causes the wave amplitude to decrease like 1/r. It is also

    follows from formulae (1.7.7), (1.7.8) for the pressure that0 p(r, t)dt = 0, even though the

    initial pressure distribution p0 > 0 within r < a.

    M. S. Howe 29 1.7 Initial value problem

  • 8/3/2019 Intro to Acoustics

    30/49

    AAS

    Figure 1.7.3

    1.7.2 Initial value problem in one dimension

    The initial value problem for one dimensional propagation of sound parallel to the x

    axis in unbounded fluid is governed by

    1c2o

    2

    t2

    2

    x2

    (pH) =

    1

    c2o

    t

    (t)f(x)

    +

    1

    c2o (t)g(x), < x < , (1.7.9)

    where f(x), g(x) are the respective initial values of p and p/t. This is solved by means

    of the Greens function (1.6.8), using the results (see equation (B.1.6) of the Appendix):1

    c2o

    ()f(y)

    G(x,y,t,)dyd =

    1

    2co

    f(y)

    t

    |x y|

    co

    dy

    =1

    2

    f(x cot) + f(x + cot)

    1c2o ()g(y)G(x,y,t,)dyd =

    1

    2co

    g(y)H

    t

    |x y|

    co

    dy

    =1

    2co

    x+cotxcot

    g(y) dy.

    The combination of these results yields DAlemberts solution

    p(x, t) =1

    2

    f(x cot) + f(x cot)

    +

    1

    2co

    x+cotxcot

    g(y) dy. (1.7.10)

    M. S. Howe 30 1.7 Initial value problem

  • 8/3/2019 Intro to Acoustics

    31/49

    AAS

    A simple application of this formula is illustrated in Figure 1.7.4, for the case where the

    fluid is initially at rest (g(x) 0) with p = f(x) = 0 only within the interval a < x < a,

    where it has the triangular waveform

    f(x) = pmax

    1

    |x|

    a

    H

    1

    |x|

    a

    with peak pressure pmax at x = 0. This pressure distribution splits symmetrically (see

    figure), forming equal waves propagating without change of form to x = , and given for

    t > 0 by

    p(x, t)

    pmax=

    1

    2

    1

    |x cot|

    a

    H

    1

    |x cot|

    a

    +

    1

    |x + cot|

    a

    H

    1

    |x + cot|

    a

    .

    Figure 1.7.4

    M. S. Howe 31 1.7 Initial value problem

  • 8/3/2019 Intro to Acoustics

    32/49

    AAS

    1.8 Monopoles, dipoles and quadrupoles

    A volume point source q(t)(x) of the type considered in 1.4 as a model for a pulsating

    sphere is also called a point monopole. For a compressible medium the corresponding

    velocity potential it produces is the outgoing solution of1

    c2o

    2

    t2 2

    = q(t)(x). (1.8.1)

    The solution can be written down by analogy with the solution (1.6.6) of equation

    (1.6.3) for the acoustic pressure. Replace p by in (1.6.6) and set F(y, ) = q()(y).

    Then

    (x, t) =q

    t |x|co

    4|x|

    q

    t rco

    4r

    . (1.8.2)

    This coincides with the corresponding solution (1.4.3) for a compact pulsating sphere.

    Changes in the motion of the sphere (i.e. in the value of the volume outflow rate q(t)) are

    communicated to a fluid element at distance r after a time delay r/co required for sound to

    travel outward from the source.

    1.8.1 The point dipole

    Let f = f(t) be a time dependent vector, then a source on the right of the acoustic

    pressure equation (1.6.3) of the form

    F(x, t) = div

    f(t)(x)

    xj

    fj(t)(x)

    (1.8.3)

    is called a point dipole (located at the origin). The repeated subscript j in this equation,

    implies a summation over j = 1, 2, 3 (see Appendix A). Equation (1.3.7) shows that

    the point dipole is equivalent to a force distribution F(x, t) = f(t)(x) per unit volume

    applied to the fluid at the origin.

    The sound produced by the dipole can be calculated from (1.6.6), but it is easier to use

    (1.6.5):

    p(x, t) =1

    4

    yj

    fj()(y)

    t |xy|co

    |x y|d3yd.

    M. S. Howe 32 1.8 Monopoles, dipoles . . .

  • 8/3/2019 Intro to Acoustics

    33/49

    AAS

    Integrate by parts with respect to each yj (recalling that (y) = 0 at yj = ), and note

    that

    yj

    t |xy|co

    |x y|

    =

    xj

    t |xy|co

    |x y|

    .

    Then

    p(x, t) =1

    4

    fj()(y)

    xj

    t |xy|co

    |x y|

    d3yd

    =1

    4

    xj

    fj()(y)

    t |xy|

    co

    |x y|

    d3yd

    i.e.

    p(x, t) = xj

    fj t |x|co

    4|x|

    . (1.8.4)The same procedure shows that for a distributed dipole source of the type F(x, t) =

    div f(x, t) on the right of equation (1.6.3), the acoustic pressure is

    p(x, t) =1

    4

    xj

    fj

    y, t |xy|co

    |x y|

    d3y. (1.8.5)

    A point dipole at the origin orientated in the direction of a unit vector n is entirely

    equivalent to two point monopoles of equal but opposite strengths placed a short distance

    apart (much smaller than the acoustic wavelength) on opposite sides of the origin on a line

    through the origin parallel to n. For example, if n is parallel to the x-axis, and the sources

    are distance apart, the two monopoles would be

    q(t)

    x

    2

    (y)(z) q(t)

    x +

    2

    (y)(z) q(t)(x)(y)(z)

    x

    q(t)(x)

    .

    This is a fluid volume dipole. The relation p = o/t implies that the equivalent dipole

    source in the pressure equation (1.3.7) or (1.6.3) is

    o

    x

    q

    t(t)(x)

    .

    M. S. Howe 33 1.8 Monopoles, dipoles . . .

  • 8/3/2019 Intro to Acoustics

    34/49

    AAS

    1.8.2 Quadrupoles

    A source distribution involving two space derivatives is equivalent to a combination of

    four monopole sources whose net volume source strength is zero, and is called a quadrupole.

    A general quadrupole is a source of the form

    F(x, t) =2Tij

    xixj(x, t) (1.8.6)

    in equation (1.6.3). The argument above leading to expression (1.8.5) can be applied twice

    to show that the corresponding acoustic pressure is given by

    p(x, t) =1

    4

    2

    xixj

    Tij(y, t |x y|/co)

    |x

    y

    |

    d3y. (1.8.7)

    1.8.3 Solution of the general linear acoustic equation

    The sources on the right of the general linear acoustic equation (1.3.7)

    1

    c2o

    2

    t2 2

    p = o

    q

    t div F

    are respectively of monopole and dipole type. The solution with outgoing wave behaviour

    is therefore

    p(x, t) =o4

    t

    q(y, t |x y|/co)

    |x y|d3y

    1

    4

    xj

    Fj(y, t |x y|/co)

    |x y|d3y.

    (1.8.8)

    1.8.4 Vibrating sphere

    Let a rigid sphere of radius a execute small amplitude oscillations at speed U(t) in

    the x1-direction (Figure 1.8.1a) at sufficiently low frequencies that it may be assumed to

    be acoustically compact. Take the coordinate origin at the mean position of the centre.

    The sphere pushes fluid away from its advancing front hemisphere, and the retreating

    rear hemisphere draws in fluid from its wake. The sphere therefore resembles a dipole

    source, and it is shown in 2.7 that the motion induced in an ideal fluid is equivalent to

    M. S. Howe 34 1.8 Monopoles, dipoles . . .

  • 8/3/2019 Intro to Acoustics

    35/49

    AAS

    that produced by a point volume dipole of strength 2a3U(t) at the position of its centre

    directed along the x1-axis. The velocity potential is the solution of

    1c2o

    2

    t2

    2 = x1

    2a3U(t)(x) . (1.8.9)By analogy with (1.8.3) and (1.8.4), we have

    (x, t) =

    x1

    2a3U

    t |x|co

    4|x|

    . (1.8.10)

    Now

    xj|x| =

    xj|x|

    . (1.8.11)

    Applying this formula for j = 1, we find (putting r = |x| and x1 = r cos )

    = a3cos

    2r2U

    t r

    co

    a3 cos

    2cor

    U

    t

    t r

    co

    near field far field

    The near field term is dominant at sufficiently small distances r from the origin that

    1r 1co1UUt fco

    where f is the characteristic frequency of the oscillations of the sphere. But, sound of

    frequency f travels a distance

    co/f = one acoustic wavelength

    in one period of oscillation 1/f. Hence the near field term is dominant when

    r .

    The motion becomes incompressible when co . In this limit the solution reduces

    entirely to the near field term, which is also called the hydrodynamic near field; it decreases

    in amplitude like 1/r2 as r .

    M. S. Howe 35 1.8 Monopoles, dipoles . . .

  • 8/3/2019 Intro to Acoustics

    36/49

    AAS

    The far field is the acoustic region that only exists when the fluid is compressible.

    It consists of propagating sound waves, carrying energy away from the sphere, and takes

    over from the near field when r . There is an intermediate zone where r in which

    the solution is in a state of transition from the near to the far field. This accords with the

    assumption that the sphere is compact and can be replaced by the point dipole: the motion

    close to the sphere is essentially the same as if the fluid is incompressible i.e. the diameter

    of the sphere is much smaller than the acoustic wavelength ( a ).

    Figure 1.8.1

    The intensity of the sound generated by the sphere in the far field is proportional to 2:

    2 a6

    4c2o

    r2 U

    t 2

    trco

    cos2 , r .

    The dependence on determines the directivity of the sound. For the dipole it has the

    figure of eight pattern illustrated in Figure 1.8.1b, with peaks in directions parallel to the

    dipole axis ( = 0, ); there are radiation nulls at = 2

    (the curve should be imagined to

    be rotated about the x1-axis).

    M. S. Howe 36 1.8 Monopoles, dipoles . . .

  • 8/3/2019 Intro to Acoustics

    37/49

    AAS

    1.9 Acoustic energy equation

    The acoustic energy equation in a stationary ideal fluid is obtained by taking the scalar

    product of the velocity v and linearised momentum equation (1.3.1). The result is written

    t

    1

    2ov

    2

    + divp v

    p div v = v F.

    The term p div v on the left is rendered in a more useful form by substitution for div v from

    the linearised equation of continuity (1.3.3)

    div v = q 1

    o

    t= q

    1

    oc2o

    p

    t.

    The required energy equation is then obtained in the form

    t

    1

    2ov

    2 +1

    2

    p2

    oc2o

    + div

    p v

    = pq + v F. (1.9.1)

    The term

    E =1

    2ov

    2 +1

    2

    p2

    oc2o, (1.9.2)

    is the acoustic energy per unit volume, the first part of which is obviously the kinetic

    energy density. The second term is the compressional or potential energy component,

    calculated as follows. Let V be the volume occupied by unit mass of the fluid, then V = 1,

    and VVo

    p dV is the compressional energy per unit mass, where Vo = 1/o is the volume

    occupied by unit mass in the absence of the sound, and p is the perturbation pressure

    produced by the volumetric change V Vo. Then the compressional energy per unit volume

    is (to second order, and using the adiabatic formula dp = c2od)

    o V

    Vop dV = o

    op d

    1

    =

    1

    oc2o p

    0p dp =

    1

    2

    p2

    oc2o.

    The divergence term in (1.9.1) governs the rate at which acoustic energy propagates

    out of unit volume of fluid. The terms on the right are respectively the rates of production

    of acoustic energy by the volume and force distribution sources. The energy balance is

    perhaps more clearly exhibited by integration of the energy equation over the fluid region V

    M. S. Howe 37 1.9 Acoustic energy

  • 8/3/2019 Intro to Acoustics

    38/49

    AAS

    bounded by a large closed surface S (Figure 1.9.1) that contains all of the acoustic sources.

    Then

    t V Ed3x + Sp

    v dS = V (pq +v F) d3x, (1.9.3)

    which equates the sum of the rate of accumulation of energy within V and the energy flux

    out through S to the rate of working of the acoustic sources.

    1.9.1 Calculation of the energy flux

    At large distances r from a source region we generally have

    p(x, t) o

    ,,t r

    co

    r

    , r , (1.9.4)

    where the function depends on the nature of the source distribution, and and are

    polar angles determining the directivity of the sound. From the radial component of the

    linearized momentum equation

    vrt

    = 1

    o

    p

    r

    1

    r2

    ,,t r

    co

    +

    1

    cor

    t

    ,,t

    r

    co

    . (1.9.5)

    The first term in the second line can be neglected when r , and therefore

    vr 1

    cor

    ,,t r

    co

    p

    oco. (1.9.6)

    By considering the and components of the momentum equation we can show that the

    corresponding velocity components v, v, say, decrease faster than 1/r as r . We

    therefore conclude from this and (1.9.6) that the acoustic particle velocity is normal to

    the acoustic wavefronts (the spherical surfaces r = cot). In other words: sound consists of

    longitudinal waves in which the fluid particles oscillate backwards and forwards along the

    local direction of propagation of the sound.

    The acoustic power radiated by the source distribution is given by the surface

    integral of equation (1.9.3), which we can take in the form

    =S

    pvr dS =S

    p2

    ocodS, (1.9.7)

    M. S. Howe 38 1.9 Acoustic energy

  • 8/3/2019 Intro to Acoustics

    39/49

    AAS

    where the surface of integration S is that of a large sphere of radius r centered on the

    source region. Because the surface area = 4r2 we only need to know the pressure and

    velocity correct to order 1/r on S in order to evaluate the integral. Smaller contributions

    (such as that determined by the first term in the second line of (1.9.5)) decrease too fast as

    r increases to supply a finite contribution to the integral as r .

    Figure 1.9.1

    In acoustic problems we are therefore usually satisfied if we can calculate the pressure

    and velocity in the acoustic far field correct to order 1/r; this will always permit the

    evaluation of the radiated sound power. The formula vr = p/oco is applicable at large

    distances from the sources, where the wavefronts can be regarded as locally plane, but it is

    true identically for plane sound waves. In the latter case, and for spherical waves on the

    surface of the large sphere of Figure 1.9.1, the quantity

    I = pvr =p2

    oco(1.9.8)

    is called the acoustic intensity. It is the rate of transmission of acoustic energy per unit

    area of wavefront. For a plane wave E = p2/oc2o, so that I = coE, i.e. the plane wave energy

    flux is equal to the energy density multiplied by the speed of sound.

    M. S. Howe 39 1.9 Acoustic energy

  • 8/3/2019 Intro to Acoustics

    40/49

    AAS

    1.9.2 Example

    For a transient acoustic source all of the wave energy radiates out through the distant

    surface S in a finite time. Integration of the energy equation (1.9.3) over all times therefore

    yields

    S

    pv dS dt =

    V

    (pq + v F) d3x dt. (1.9.9)

    Let us verify this formula for the initial value problem of 1.7.1 of sound produced by the

    sudden release of high pressure air from within the spherical region |x| = r < a.

    Outside the source region the pressure is given by equation (1.7.7), and the power

    radiated through S is

    =S

    p2

    ocodS = 4r2

    p2o

    4r2oco(cot r)

    2

    H(cot r + a) H(cot r a)

    .

    The total radiated acoustic energy E, say, determined by the left hand side of (1.9.9) is

    therefore

    E =p20oco

    (r+a)/co(ra)/co

    (cot r)2 dt =

    2a3

    3

    p20oc2o

    . (1.9.10)

    On the right of (1.9.9) we have, from equations (1.7.5), F = 0 and

    q =p0

    oc2o(t)H(a r).

    This must be multiplied by the pressure given by equation (1.7.8) for r < a, following which

    E is calculated by evaluation of E =

    V pq d

    3x dt. But in doing this it must be recalled

    that the radiated pressure is the causal solution of equation (1.7.1), and what is actually

    calculated is not p but pH(t). Thus,

    E = p2

    0

    oc2o

    |x|

  • 8/3/2019 Intro to Acoustics

    41/49

    AAS

    which agrees with (1.9.10).

    The integration with respect to time has made use of the result dH(t)/dt = (t). Readers

    uncomfortable with this formal step should observe that the solution of the initial value

    problem (1.7.1) can be carried through with negligible change of details, by replacing H( t)

    and (t) by their corresponding sequences H(t), (t) (see Appendix B)

    H(t) =1

    2+

    1

    tan1

    t

    , (t) =

    (2 + t2), > 0,

    in which case H(t) (t) dt = [

    12H

    2(t)]

    =

    12 .

    M. S. Howe 41 1.9 Acoustic energy

  • 8/3/2019 Intro to Acoustics

    42/49

    AAS

    1.10 Calculation of the acoustic far field

    We now discuss the approximations necessary to evaluate the sound in the far field from

    the retarded potential representation:

    p(x, t) =1

    4

    F

    y, t |xy|co

    |x y|

    d3y. (1.10.1)

    We assume that F(x, t) = 0 only within a finite source region (Figure 1.10.1), and take the

    coordinate origin O within the region.

    Figure 1.10.1

    When |x| and y lies within the source region (so that |x| |y|)

    |x y|

    |x|2 2x y + |y|212

    = |x|

    1 2x y

    |x|2 +|y|2

    |x|2 12

    |x|

    1

    x y

    |x|2+ O

    |y|2

    |x|2

    i.e. |x y| |x| x y

    |x|when

    |y|

    |x| 1. (1.10.2)

    Similarly,1

    |x y|

    1

    |x| xy|x|

    1

    |x|

    1 +

    x y

    |x|2

    . .1

    |x y|

    1

    |x|+

    x y

    |x|3when

    |y|

    |x| 1. (1.10.3)

    The approximation (1.10.3) shows that, in order to obtain the far field approximation

    of the solution (1.10.1) that behaves like 1/r = 1/|x| as |x| , it is sufficient to replace

    M. S. Howe 42 1.10 Acoustic far field

  • 8/3/2019 Intro to Acoustics

    43/49

    AAS

    |x y| in the denominator of the integrand by |x|. However, in the argument of the source

    strength F it is important to retain possible phase differences between the sound waves

    generated by components of the source distribution at different locations y; we therefore

    replace |x y| in the retarded time by the right hand side of (1.10.2). Hence,

    p(x, t) 1

    4|x|

    F

    y, t

    |x|

    co+

    x.y

    co|x|

    d3y, |x| . (1.10.4)

    This is called the Fraunhofer approximation.

    The source region may extend over many characteristic acoustic wavelengths of the

    sound. By retaining the contribution x y/co|x| to the retarded time we ensure that any

    interference between waves generated at different positions within the source region iscorrectly described by the far field approximation. In Figure 1.10.1 the acoustic travel

    time from a source point y to the far field point x is equal to that from the point labelled

    A to x when |x| . The travel time over the distance OA is just x y/co|x|, so that

    |x|/co x y/co|x| gives the correct value of the retarded time when |x| .

    1.10.1 Dipole source distributions

    By applying the far field formula (1.10.4) to a dipole source F(x, t) = div f(x, t) we

    obtain (from (1.8.5))

    p(x, t) 1

    4

    xj

    1

    |x|

    fj

    y, t

    |x|

    co+

    x y

    co|x|

    d3y

    1

    4|x|

    xj

    fj

    y, t

    |x|

    co+

    x y

    co|x|

    d3y, |x| , (1.10.5)

    because the differential operator /xj need not be applied to 1/|x| as this would give a

    contribution decreasing like 1/r2 at large distances from the dipole.

    However, it is useful to make a further transformation that replaces /xj by the time

    derivative /t, which is usually more easily estimated in applications. To do this we

    M. S. Howe 43 1.10 Acoustic far field

  • 8/3/2019 Intro to Acoustics

    44/49

    AAS

    observe that

    fjxj

    y, t

    |x|

    co+

    x y

    co|x|

    =

    fjt

    y, t

    |x|

    co+

    x y

    co|x|

    xj

    t

    |x|

    co+

    x y

    co|x|

    =fjt

    y, t

    |x|

    co+

    x y

    co|x|

    xjco|x|

    +yj

    co|x|

    (x y)xjco|x|3

    xj

    co|x|

    fjt

    y, t

    |x|

    co+

    x y

    co|x|

    as |x| .

    Hence, the far field of a distribution of dipoles F(x, t) = div f(x, t) is given by

    p(x, t) =xj

    4co|x|2

    t

    fj y, t |x|

    co+

    x y

    co|x| d3y. (1.10.6)

    Note thatxj

    |x|2=

    xj|x|

    1

    |x|,

    where xj/|x| is the jth component of the unit vector x/|x|. Thus, the additional factor of

    xj/|x| in (1.10.6) does not change the rate of decay of the sound with distance from the

    source (which is still like 1/r) but it does have an influence on the acoustic directivity.

    A comparison of (1.10.5) and (1.10.6) leads to the following rule for interchanging space

    and time derivatives in the acoustic far field

    xj

    1

    co

    xj|x|

    t. (1.10.7)

    1.10.2 Quadrupole source distributions

    For the quadrupole (1.10.6)

    F(x, t) =2Tij

    xixj(x, t)

    and

    p(x, t) =1

    4

    2

    xixj

    Tij(y, t |x y|/co)

    |x y|d3y.

    M. S. Howe 44 1.10 Acoustic far field

  • 8/3/2019 Intro to Acoustics

    45/49

    AAS

    By applying (1.10.4) and the rule (1.10.7), we find that the acoustic far field is given by

    p(x, t) xixj

    4c2o|x|3

    2

    t2

    Tij

    y, t

    |x|

    co+

    x y

    co|x|

    d3y, |x| . (1.10.8)

    1.10.3 Example

    For the (1,2) point quadrupole

    F(x, t) =2

    x1x2(T(t)(x))

    equation (1.10.8) shows that in the acoustic far field

    p(x, t) x1x2

    4c2o|x|3

    2T

    t2

    t |x|

    co

    , |x| .

    If we use spherical polar coordinates, such that

    x1 = r cos , x2 = r sin cos , x3 = r sin sin ,

    we can write the pressure in the form

    p(x, t) sin2 cos

    8c2o|x|

    2T

    t2

    t

    |x|

    co

    , |x| .

    The directivity of the sound ( p2) is therefore represented by sin2 2 cos2 . Its shape is

    plotted in Figure 1.10.2 for radiation in the x1x2-plane ( = 0, ). The four lobe clover

    leaf pattern is characteristic of a quadrupole Tij for which i = j.

    M. S. Howe 45 1.10 Acoustic far field

  • 8/3/2019 Intro to Acoustics

    46/49

    AAS

    Figure 1.10.2

    1.10.4 Far field of a compact source distribution

    Phase variations x y/co|x| of sound arriving from different parts of a generalised source

    F(x, t) are small if the source region is acoustically compact. Therefore, the far field

    approximation (|x| ) (1.10.4) becomes

    p(x, t) 1

    4|x|

    F

    y, t

    |x|

    co+

    x.y

    co|x|

    d3y

    1

    4|x|

    F

    y, t

    |x|

    co

    d3y.

    This gives the principal component of the radiating sound unless it should happen that the

    overall source strength is null, i.e., unless F(y, t) d

    3y = 0. The amplitude of sound

    waves in the far field is then crucially dependent on the existence of small phase mismatches

    between different parts of the source, and must be determined by expanding the Fraunhofer

    approximation in powers of x y/co|x|:

    M. S. Howe 46 1.10 Acoustic far field

  • 8/3/2019 Intro to Acoustics

    47/49

    AAS

    p(x, t) xi

    4co|x|2

    t

    yiF

    y, t

    |x|

    co

    d3y +

    xixj8co|x|3

    2

    t2

    yiyjF

    y, t

    |x|

    co

    d3y

    +

    = dipole + quadrupole + (1.10.9)

    Each term in this multipole expansion is nominally of order /co = ko 1 relative to

    its predecessor, where /t is the characteristic source frequency and the diameter

    of the source region. Therefore, the expansion is halted at the first non-zero term (see

    Question 9 of Problems 1).

    M. S. Howe 47 1.10 Acoustic far field

  • 8/3/2019 Intro to Acoustics

    48/49

    AAS

    Problems 1

    1. Derive the relations (1.3.12) for plane sound waves.

    2. Use the trial solution = (t |x|/co) to solve the problem1

    c2o

    2

    t2

    2

    x2

    = (x)(t), < x < ,

    where = 0 for t < 0.

    3. When the body force F = g in equation (1.3.1), where g is the acceleration due to gravity,

    show that in the adiabatic approximation the linearised acoustic wave equation becomes

    1

    oc2o

    2p

    t2

    xj 1

    o

    p

    xj = q

    t div pg

    oc2o ,

    where p = p po, and the mean pressure, density and sound speed po, o, co vary with depth

    in the atmosphere. Deduce that the gravitational term on the right hand side can be neglected

    provided g/co, where is a typical acoustic frequency.

    4. Consider the initial value problem (1.7.1), (1.7.2) in which the pressure p = p0 = constant and

    p/t = 0 at time t = 0 in 0 < x1 < h. Show that Poissons solution (1.7.4) predicts plane wave

    propagation parallel to the x1 direction, and that the pressure wave radiating through a typical

    exterior point P distance r1 from the source region has amplitude12p0, arrives at time t = r1/co

    and occupies a plane shell of thickness h.

    M. S. Howe 48 Problems 1

  • 8/3/2019 Intro to Acoustics

    49/49

    AAS

    5. Derive DAlemberts solution (1.7.10) for the initial value problem in one space dimension from

    Poissons formula (1.7.4).

    6. At time t = 0 a plane acoustic wave p = p(x cot) propagating in unbounded fluid satisfies

    p = p(x), p/t = cop(x) (the prime denoting differentiation with respect to the argument).

    Verify that Poissons solution (1.7.4) predicts that p = p(x cot) when t > 0.

    7. Calculate the acoustic power (1.9.7) radiated by an acoustically compact sphere of radius a exe-

    cuting small amplitude translational oscillations of frequency and velocity U(t) = Uo cos(t),

    where Uo = constant.

    8. As for Problem 7 when the sphere executes small amplitude radial oscillations at normal velocity

    vn = Uo cos(t), Uo = constant.

    9. Consider the outgoing wave solution of

    1

    c2o

    2

    t22

    =

    sgn(x)H(a |x|)U(t)

    a, a > 0,

    where U(t) is known as a function of time. If the source region |x| < a is acoustically compact,

    show that

    (x, t) a3 cos

    8co|x|

    U

    tt

    |x|

    co , |x| ,

    where cos = x/|x|.

    10. A volume point source of strength qo(t) translates at constant, subsonic velocity U. The velocity

    potential (x, t) of the radiated sound is determined by the solution of

    1

    c2o

    2

    t22

    = qo(t)(xUt).

    Show that

    (x, t) =

    qo(tR/co)

    4R(1Mcos ), M =

    U

    co ,

    where R is the distance of the reception pointx from the source position at the time of emission of

    the sound received at x at time t, and is the angle between U and the direction of propagation


Recommended