2
Waves in Fluids

2.1 Introduction

The acoustic wave motion is described by the equations of aerodynamics that are linearized because of the small fluctuations that occur in acoustic waves compared to the static state variables. The fluid motion is described generally by three equations:

  • Continuity equation – conservation of mass
  • Newton’s law – conservation of momentum
  • State law – pressure volume relationship.

For lumped systems the velocity is simply the time derivative of the point mass position in space. The same approach can be used in fluid dynamics, but here the continuous fluid is subdivided into several cells and their movement is described by trajectories. This is called the Lagrange description of fluid dynamics. Even if the equation of motions are simpler in that formulation it is quite complicated to follow all coordinates of fluid volumes in a complex flow. Thus, the Euler description of fluid dynamics is used. In this description the conservation equations are performed for a control volume that is fixed in space and the flow passes through this volume.

In this chapter, the three dimensional space is given by the Cartesian coordinates x={x,y,z}T and the velocity of the fluid v={ vx,vy,vz }T.

2.2 Wave Equation for Fluids

2.2.1 Conservation of Mass

For simplicity we consider first the flow in the x-direction as in Figure 2.1. The mass flow balance contains the following quantities:

  1. The elemental mass m=ρdV=ρA with A=dydz.
  2. Mass flow into the volume (ρvxA)x.
  3. Mass flow out of the volume (ρvxA)x+dx.
  4. Mass input from external sources m˙.

Figure 2.1Mass flow in x-direction through control volume.
Source: Alexander Peiffer.

leading to equation

StartFraction partial-differential left-parenthesis rho upper A d x right-parenthesis Over partial-differential t EndFraction equals left-parenthesis rho v Subscript normal x Baseline upper A right-parenthesis Subscript x Baseline minus left-parenthesis rho v Subscript normal x Baseline upper A right-parenthesis Subscript x plus d x Baseline plus ModifyingAbove m With dot  (2.1)

for mass conservation. Expanding the second term on the right hand side in a Taylor series gives

StartFraction partial-differential left-parenthesis rho upper A d x right-parenthesis Over partial-differential t EndFraction equals left-bracket left-parenthesis rho v Subscript normal x Baseline upper A right-parenthesis Subscript x Baseline minus left-parenthesis rho v Subscript normal x Baseline upper A right-parenthesis Subscript x Baseline minus StartFraction partial-differential left-parenthesis rho v Subscript normal x Baseline upper A right-parenthesis Subscript x Baseline Over partial-differential x EndFraction d x right-bracket plus ModifyingAbove m With dot

and finally

StartFraction partial-differential rho Over partial-differential t EndFraction plus StartFraction partial-differential left-parenthesis rho v Subscript normal x Baseline right-parenthesis Over partial-differential x EndFraction equals ModifyingAbove rho With dot Subscript normal s  (2.2)

This one dimensional equation of mass conservation in x-direction can be extended to three dimensions:

StartFraction partial-differential rho Over partial-differential t EndFraction plus nabla left-parenthesis rho bold v right-parenthesis equals ModifyingAbove rho With dot Subscript normal s  (2.3)

The second term of (2.3) may be confusing, but it says that the change of density is not only determined by a gradient in the velocity field but also by a gradient of the density.

2.2.2 Newton’s law – Conservation of Momentum

The same procedure is applied to the momentum of the fluid. As shown in Figure 2.2 we get for flow in x-direction:

  1. The momentum of the control volume is ρvxdV=ρvxAdx.
  2. The momentum flow into the volume (ρvx2A)x.
  3. mass flow out of the volume (ρvx2A)x+dx.
  4. The force at position x is (PA)x.
  5. The force at position x+dx is (PA)x+dx.
  6. External volume force density fx.

Figure 2.2Momentum flow in x-direction through control volume.
Source: Alexander Peiffer.

Thus, the conservation of momentum in x reads

StartFraction partial-differential left-parenthesis rho v Subscript normal x Baseline upper A d x right-parenthesis Over partial-differential t EndFraction equals left-parenthesis rho v Subscript normal x Superscript 2 Baseline upper A right-parenthesis Subscript x Baseline minus left-parenthesis rho v Subscript normal x Superscript 2 Baseline upper A right-parenthesis Subscript x plus d x Baseline plus left-parenthesis upper P upper A right-parenthesis Subscript x Baseline minus left-parenthesis upper P upper A right-parenthesis Subscript x plus d x Baseline plus upper F Subscript x  (2.4)

Using Taylor expansions for (ρvx2A)x+dx and (PA)x+dx gives

StartFraction partial-differential left-parenthesis rho v Subscript normal x Baseline right-parenthesis Over partial-differential t EndFraction equals minus StartFraction partial-differential left-parenthesis rho u Subscript x Superscript 2 Baseline right-parenthesis Over partial-differential x EndFraction minus StartFraction partial-differential upper P Over partial-differential x EndFraction plus f Subscript x  (2.5)

Here, fx=Fx/(Adx) is the volume force density (force per volume). Using the chain law the partial derivatives of the first and second term lead to

rho StartFraction partial-differential v Subscript normal x Baseline Over partial-differential t EndFraction plus v Subscript normal x Baseline left-brace StartFraction partial-differential rho Over partial-differential t EndFraction plus v Subscript normal x Baseline StartFraction partial-differential rho Over partial-differential x EndFraction plus rho StartFraction partial-differential v Subscript normal x Baseline Over partial-differential x EndFraction right-brace plus rho v Subscript normal x Baseline StartFraction partial-differential v Subscript normal x Baseline Over partial-differential x EndFraction plus StartFraction partial-differential upper P Over partial-differential x EndFraction equals f Subscript x  (2.6)

The term in brackets is the homogeneous continuity Equation (2.1), and Equation (2.6) simplifies to

rho StartFraction partial-differential v Subscript normal x Baseline Over partial-differential t EndFraction plus rho v Subscript normal x Baseline StartFraction partial-differential v Subscript normal x Baseline Over partial-differential x EndFraction plus StartFraction partial-differential upper P Over partial-differential x EndFraction equals f Subscript x  (2.7)

As with the conservation of mass, this can be extended to three dimensions:

rho left-brace StartFraction partial-differential bold v Over partial-differential t EndFraction plus left-parenthesis bold v nabla right-parenthesis bold v right-brace plus nabla upper P equals bold f  (2.8)

This equation is the non-linear, inviscid momentum equation called the Euler equation.

2.2.3 Equation of State

The above equations relate pressure, velocity and density. For further reducing this set we need a third equation. The easiest way would be to introduce the . Here we start with the first law of thermodynamics in order to show the difference between isotropic (or adiabatic) equation of state and other relationships.

d q equals d u plus upper P d v minus d r  (2.9)

With the following specific quantities per unit mass

StartLayout 1st Row 1st Column d q 2nd Column period period period 3rd Column specific heat q equals q left-parenthesis upper T comma rho right-parenthesis 2nd Row 1st Column d v 2nd Column period period period 3rd Column specific volume v equals upper V slash upper M 3rd Row 1st Column upper P d v 2nd Column period period period 3rd Column specific expansion work 4th Row 1st Column d r 2nd Column period period period 3rd Column specific friction losses EndLayout  (2.10)

With the specific entropy ds=dq+drT we get:

StartLayout 1st Row 1st Column d s equals left-parenthesis StartFraction partial-differential u Over partial-differential upper T EndFraction right-parenthesis Subscript upper T Baseline StartFraction d upper T Over upper T EndFraction 2nd Column plus 3rd Column StartFraction upper P Over upper T EndFraction d upper V 2nd Row 1st Column equals StartFraction c Subscript normal v Baseline Over upper T EndFraction d upper T 2nd Column plus 3rd Column StartFraction upper P Over upper T EndFraction d upper V 3rd Row 1st Column equals StartFraction c Subscript normal v Baseline Over upper T EndFraction d upper T 2nd Column plus 3rd Column StartFraction upper P Over upper T EndFraction d left-parenthesis StartFraction 1 Over rho EndFraction right-parenthesis 4th Row 1st Column equals StartFraction c Subscript normal v Baseline Over upper T EndFraction d upper T 2nd Column plus 3rd Column StartFraction upper P Over upper T rho squared EndFraction d rho EndLayout  (2.10)

The relation dv=d(1/ρ) comes from the fact that v is a mass specific value and therefore the reciprocal of the density ρ=1/v. For an ideal gas we have

upper P equals rho upper R upper T with upper R equals c Subscript normal p Baseline minus c Subscript normal v Baseline  (2.11)

cp and cv are the specific thermal heat capacities for constant pressure and volume, respectively. That is the ratio of temperature change T per increase of heat q. From the total differential

d upper P left-parenthesis upper T comma rho right-parenthesis equals left-parenthesis StartFraction partial-differential p Over partial-differential upper T EndFraction right-parenthesis Subscript rho Baseline d upper T plus left-parenthesis StartFraction partial-differential p Over partial-differential rho EndFraction right-parenthesis Subscript upper T Baseline d rho  (2.12)

we can derive

StartFraction d upper T Over upper T EndFraction equals StartFraction d upper P Over upper P EndFraction minus StartFraction d rho Over rho EndFraction  (2.13)

Using all above relations the change in density dρ is:

d rho equals StartFraction rho Over kappa upper P EndFraction d upper P minus StartFraction rho Over c Subscript normal p Baseline EndFraction d s  (2.14)

with κ=cv/cp. In most acoustic cases the process is isotropic: i.e. time scales are too short for heat exchange in a free gas; thus ds=0, and the change of pressure per density is

left-parenthesis StartFraction d upper P Over d rho EndFraction right-parenthesis Subscript normal s Baseline equals kappa StartFraction upper P Over rho EndFraction equals c 0 squared  (2.15)

In case of constant temperature (isothermal) dT=0 we get with (2.12) and the ideal gas law (2.11):

left-parenthesis StartFraction d upper P Over d rho EndFraction right-parenthesis Subscript upper T Baseline equals StartFraction upper P Over rho EndFraction equals c Subscript 0 upper T Superscript 2  (2.16)

As we will later see, c0 is the . Newton calculated the wrong speed of sound based on the assumption of constant temperature that was later corrected by Laplace by the conclusion that the process is adiabatic. For fluids and liquids like water a different quantity is used because there is no such expression as the ideal gas law. The bulk modulus is defined as:

upper K equals rho left-parenthesis StartFraction partial-differential upper P Over partial-differential rho EndFraction right-parenthesis  (2.17)

Due to (2.15) and (2.16) the relationship between the bulk modulus K and c0 is:

c 0 squared equals StartFraction upper K Over rho EndFraction  (2.18)

The bulk modulus can be defined for gases too, but we must distinguish between isothermal or adiabatic processes.

StartLayout 1st Row 1st Column upper K Subscript normal s 2nd Column rho left-parenthesis StartFraction partial-differential upper P Over partial-differential rho EndFraction right-parenthesis Subscript normal s Baseline equals kappa upper P 3rd Column upper K Subscript upper T 4th Column rho left-parenthesis StartFraction partial-differential upper P Over partial-differential rho EndFraction right-parenthesis Subscript upper T Baseline equals upper P EndLayout  (2.19)

2.2.4 Linearized Equations

Equations (2.3) and (2.8) can be linearized if small changes around a certain equilibrium are considered:

StartLayout 1st Row 1st Column rho 2nd Column equals 3rd Column rho 0 plus rho prime EndLayout  (2.20)
StartLayout 1st Row 1st Column upper P 2nd Column equals 3rd Column upper P 0 plus p EndLayout  (2.21)
StartLayout 1st Row 1st Column bold v 2nd Column equals 3rd Column bold v 0 plus bold v prime EndLayout  (2.22)

Inserting (2.22) into the equation of continuity (2.3), neglecting all second order terms as far as source terms, and setting1 v0=0 the linear equation of continuity is:

StartFraction partial-differential rho prime Over partial-differential t EndFraction plus rho 0 nabla bold v Superscript prime Baseline equals 0  (2.23)

Doing the same for the equation of motion (2.8) leads to:

rho 0 StartFraction partial-differential bold v prime Over partial-differential t EndFraction plus nabla p equals 0  (2.24)

Using the curl(×) of this equation it can be shown that the acoustic velocity v can be expressed using a so-called velocity potential which will be useful for the calculation of some wave propagation phenomena.

bold v prime equals nabla normal upper Phi  (2.25)

2.2.5Acoustic Wave Equation

From the following operation

StartFraction partial-differential Over partial-differential t EndFraction left-parenthesis 2.23 right-parenthesis minus nabla left-parenthesis 2.24 right-parenthesis

follows

StartFraction partial-differential squared rho prime Over partial-differential t squared EndFraction minus nabla squared p equals 0  (2.26)

With the equation of state (2.15) for the density we get the linear wave equation for the acoustic pressure p

StartFraction 1 Over c 0 squared EndFraction StartFraction partial-differential squared p Over partial-differential t squared EndFraction minus nabla squared p equals 0  (2.27)

Inserting the velocity v=Φ derived from the potential Φ into the linear equation of motion (2.24) provides the required relation between pressure and the velocity potential

rho 0 StartFraction partial-differential Over partial-differential t EndFraction nabla normal upper Phi plus nabla p equals nabla left-parenthesis rho 0 StartFraction partial-differential normal upper Phi Over partial-differential t EndFraction plus p right-parenthesis equals 0  (2.28)

Thus, the relationship between pressure p and the velocity potential Φ is

p equals minus rho 0 StartFraction partial-differential normal upper Phi Over partial-differential t EndFraction  (2.29)

Entering this into the wave equation (2.27) and eliminating one time derivative gives:

StartFraction 1 Over c 0 squared EndFraction StartFraction partial-differential squared normal upper Phi Over partial-differential t squared EndFraction minus nabla squared normal upper Phi equals 0  (2.30)

The definition of the velocity potential (2.25) and equation (2.29) can be applied for the derivation of a relationship between acoustic velocity and pressure:

bold v Superscript prime Baseline equals nabla normal upper Phi equals minus StartFraction 1 Over rho 0 EndFraction integral nabla p d t  (2.31)

2.3 Solutions of the Wave Equation

In acoustics we stay in most cases in the linear domain, so we change the notations from equations (2.20)–(2.22):

bold v Superscript prime Baseline right-arrow bold v rho Superscript prime Baseline right-arrow rho  (2.32)

Equations (2.27) and (2.30) define the mathematical law for the propagation of waves. For the explanation of basic concepts the wave equation is used in one dimensional form.

2.3.1 Harmonic Waves

According to D’Alambert every function of the form p(x,t)=Af(xc0t)+Bg(x+c0t) is a solution of the one-dimensional wave equation. In the following we will consider harmonic motion or waves so we replace the functions f and g by the exponential function with

f left-parenthesis x right-parenthesis comma g left-parenthesis x right-parenthesis equals e Superscript j omega x slash c 0 Baseline equals e Superscript j k x Baseline with k equals StartFraction omega Over c 0 EndFraction  (2.33)

and get

bold-italic p left-parenthesis x comma t right-parenthesis equals bold-italic upper A e Superscript j left-parenthesis omega t minus k x right-parenthesis Baseline plus bold-italic upper B e Superscript j left-parenthesis omega t plus k x right-parenthesis  (2.34)

The first term of the right hand side of this equation is travelling in positive directions, the second in negative directions2. Harmonic waves are characterized by two quantities, the angular frequency ω and the wavenumber k. The first is the frequency (in time) as for the harmonic oscillator, and the second is a frequency in space. A similar relationship can be found between the time period T and the wavelength λ. Space and time domains are coupled by the sound velocity c0 as shown in Table 2.1.

Table 2.1 Quantities of wave propagation in time and space domains.

NameTimeSpace
SymbolUnitSymbolUnit
PeriodTsλ=c0Tm
Frequencyf=1Ts 1(Hz)()=1λm 1
Angular frequencyω=2πf=2πTs 1k=2πλ=ω/c0m 1

The time integration in Equation (2.31) corresponds to the factor 1/(jω) and reads in the frequency domain:

bold v equals minus StartFraction 1 Over j omega rho 0 EndFraction nabla bold-italic p  (2.35)

For one-dimensional waves in the x-direction this leads to:

bold-italic v Subscript x Baseline equals minus StartFraction 1 Over j omega rho 0 EndFraction StartFraction partial-differential bold-italic p Over partial-differential x EndFraction equals StartFraction 1 Over rho 0 c 0 EndFraction left-parenthesis bold-italic upper A e Superscript j left-parenthesis omega t minus k x right-parenthesis Baseline minus bold-italic upper B e Superscript j left-parenthesis omega t plus k x right-parenthesis Baseline right-parenthesis  (2.36)

Depending on the wave orientation the ratio between pressure and velocity is given by:

bold-italic v Subscript x Baseline equals plus-or-minus StartFraction 1 Over rho 0 c 0 EndFraction bold-italic p  (2.37)

In accordance with the impedance concept from section 1.2.3 we define the ratio of complex pressure and velocity as specific acoustic impedance z

bold-italic z equals StartFraction bold-italic p Over bold-italic v EndFraction  (2.38)

also called acoustic impedance. For plane waves this leads to:

z 0 equals plus-or-minus rho 0 c 0  (2.39)

Figure 2.3One-dimensional harmonic waves travelling in the positive x-direction (c0=2m/s, T=2.2s).
Source: Alexander Peiffer.

z0=ρ0c0 is called the characteristic acoustic impedance of the fluid. The specific acoustic impedance z is complex, because for waves that are not plane the velocity may be out of phase with the pressure. However, for plane waves the specific acoustic impedance is real and an important fluid property.

The above description of plane waves can be extended to three-dimensional space by introducing a wavenumber vector k.

bold-italic p equals bold-italic upper A e Superscript j left-parenthesis omega t minus bold k bold x right-parenthesis  (2.40)

2.3.2 Helmholtz equation

Entering (2.40) into the wave Equation (2.27) provides

left-parenthesis StartFraction 1 Over c 0 squared EndFraction StartFraction partial-differential squared Over partial-differential t squared EndFraction minus normal upper Delta right-parenthesis bold-italic p left-parenthesis bold x comma t right-parenthesis equals minus left-parenthesis StartFraction omega squared Over c 0 squared EndFraction plus normal upper Delta right-parenthesis bold-italic p left-parenthesis bold x comma omega right-parenthesis e Superscript j omega t Baseline equals 0 period  (2.41)

The ejωt term is often omitted and with k=ω/c0 we get the homogeneous.

left-parenthesis k squared plus normal upper Delta right-parenthesis bold-italic p left-parenthesis bold x comma omega right-parenthesis equals 0  (2.42)

2.3.3 Field Quantities: Sound Intensity, Energy Density and Sound Power

A sound wave carries a certain amount of energy that is moving with the speed of sound. We start with the instantaneous acoustic power Π:

normal upper Pi left-parenthesis t right-parenthesis equals bold upper F bold v  (2.43)

F is the force acting on a fluid particle and v the associated velocity. The acoustic intensity I is defined as the power per unit area A=An in the direction of the unit vector n and with F=pAn we get:

bold upper I left-parenthesis t right-parenthesis equals p bold v bold n  (2.44)

As in Equation (1.48) the time average is given by:

mathematical left-angle bold upper I mathematical right-angle Subscript upper T Baseline equals StartFraction 1 Over upper T EndFraction integral Subscript 0 Superscript upper T Baseline p bold v d t right double arrow one-half upper R e left-bracket bold-italic p bold-italic v Subscript x Superscript asterisk Baseline right-bracket  (2.45)

Using the harmonic plane wave solutions for pressure (2.34) and velocity (2.36)

StartLayout 1st Row 1st Column bold-italic p left-parenthesis x comma t right-parenthesis 2nd Column equals 3rd Column bold-italic upper A e Superscript j left-parenthesis omega t minus k x right-parenthesis 2nd Row 1st Column bold-italic v Subscript x Superscript asterisk Baseline left-parenthesis x comma t right-parenthesis 2nd Column equals 3rd Column StartFraction bold-italic upper A Superscript asterisk Baseline Over rho 0 c 0 EndFraction e Superscript j left-parenthesis omega t minus k x right-parenthesis EndLayout

the time averaged mean intensity yields:

mathematical left-angle upper I mathematical right-angle Subscript upper T Baseline equals StartFraction 1 Over upper T EndFraction integral Subscript 0 Superscript upper T Baseline StartFraction StartAbsoluteValue upper A EndAbsoluteValue squared Over 2 rho 0 c 0 EndFraction upper R e left-bracket e Superscript j Baseline 2 left-parenthesis omega t minus k x right-parenthesis Baseline right-bracket equals StartFraction 1 Over upper T EndFraction integral Subscript 0 Superscript upper T Baseline StartFraction ModifyingAbove p With caret squared Over 2 rho 0 c 0 EndFraction cosine squared left-parenthesis omega t minus k x right-parenthesis  (2.46)

and finally:

mathematical left-angle upper I mathematical right-angle Subscript upper T Baseline equals StartFraction ModifyingAbove p With caret squared Over 2 rho 0 c 0 EndFraction equals StartFraction p Subscript normal r normal m normal s Superscript 2 Baseline Over rho 0 c 0 EndFraction equals one-half rho 0 c 0 ModifyingAbove v With caret squared  (2.47)

We see that the specific impedance z0=ρ0c0 relates the intensity to the squared pressure.

The kinetic energy density ekin in a control volume V0 is written as

e Subscript normal k normal i normal n Baseline equals StartFraction upper E Subscript normal k normal i normal n Baseline Over upper V 0 EndFraction equals one-half rho 0 v Subscript normal x Superscript 2 Baseline equals StartFraction p squared Over 2 rho 0 c 0 squared EndFraction  (2.48)

The potential energy density epot follows from the adiabatic work integral as in equation (2.9)

e Subscript normal p normal o normal t Baseline equals StartFraction upper E Subscript normal p normal o normal t Baseline Over upper V 0 EndFraction equals minus StartFraction 1 Over upper V 0 EndFraction integral Subscript upper V Baseline 0 Superscript upper V Baseline upper P d upper V  (2.49)

If we use Equation (2.15) we get the change in density as a start for the change in volume

d rho equals StartFraction 1 Over c 0 squared EndFraction d upper P

With unit mass M in the control volume V0 it follows from ρ=M/V0 that

d upper V equals minus StartFraction upper V Over rho 0 EndFraction d rho equals minus StartFraction upper V Over rho 0 c 0 squared EndFraction d upper P

Finally we get:

e Subscript normal p normal o normal t Baseline equals StartFraction upper E Subscript normal p normal o normal t Baseline Over upper V 0 EndFraction equals integral Subscript upper P 0 Superscript upper P 0 plus p Baseline StartFraction upper P Over rho 0 c 0 squared EndFraction d upper P equals StartFraction p squared Over 2 rho 0 c 0 squared EndFraction  (2.50)

Pressure and velocity are in phase for plane waves; the same is true for the potential and kinetic energy density, so the total energy density is given by:

e left-parenthesis x comma t right-parenthesis equals e Subscript normal k normal i normal n Baseline plus e Subscript normal p normal o normal t Baseline equals StartFraction p left-parenthesis x comma t right-parenthesis squared Over rho 0 c 0 squared EndFraction  (2.51)

Using Equation (2.34) the time average over one period leads to:

mathematical left-angle e mathematical right-angle Subscript upper T Baseline equals StartFraction ModifyingAbove p With caret squared Over 2 rho 0 c 0 squared EndFraction equals StartFraction p Subscript normal r normal m normal s Superscript 2 Baseline Over rho 0 c 0 squared EndFraction  (2.52)

Finally, we can see that the speed of sound relates energy density to the sound intensity.

e equals StartFraction upper I Over c 0 EndFraction  (2.53)

All those above expressions are useful for the description and evaluation of sound fields. Especially in case of statistical methods that are based on the energy density of acoustic subsystems they link the wave fields to the energy in the systems and the power irradiating at the system boundaries.

If the intensity can be determined over a certain surface the source power is calculated by integrating the intensity component perpendicular to the surface

normal upper Pi equals integral Underscript upper A Endscripts bold upper I d bold upper A  (2.54)

2.3.4 Damping in Waves

There is no motion without damping, and a sound wave propagating over a long distance will vanish. This is considered by adding a damping component to the one-dimensional solution of the wave equation similar to the decay rate in (1.22)

bold-italic p equals bold-italic upper A e Superscript minus alpha x Baseline e Superscript j left-parenthesis omega t minus k x right-parenthesis  (2.55)

Here α is the damping constant. There are several reasons for the attenuation of acoustic waves:

  • Viscous damping due to inner viscosity.
  • Thermal damping due to irreversible heat flow during wave propagation.
  • Molecular damping due to excitation of degrees of freedom of molecules (for additional content of the gas, e.g. humidity in air).

The damping loss η as defined in (1.68) is based on the amount of energy dissipated during one cycle of wave motion. The harmonic pressure wave performs one cycle of oscillation in one period in time T or space λ. So we get for η:

eta equals StartFraction 1 Over 2 pi EndFraction StartFraction normal upper Delta upper E Over upper E EndFraction equals StartFraction 1 Over 2 pi EndFraction StartFraction bold-italic upper A squared minus bold-italic upper A squared e Superscript minus 2 alpha lamda Baseline Over bold-italic upper A squared EndFraction  (2.56)

For small damping the exponential function can be approximated by ex1+x providing the relationship between damping loss and fluid wave attenuation.

eta almost-equals StartFraction 1 Over 2 pi EndFraction 2 alpha lamda with lamda equals StartFraction 2 pi Over k EndFraction  (2.57)

Hence, the attenuation can be given by:

alpha equals eta StartFraction k Over 2 EndFraction eta equals StartFraction 2 alpha Over k EndFraction  (2.58)

An appropriate way to consider this relationship in the solution of the wave equation is to include this into a complex wavenumber k:

bold-italic p equals bold-italic upper A e Superscript j left-parenthesis minus bold-italic k x plus omega t right-parenthesis Baseline equals bold-italic upper A e Superscript minus StartFraction eta x Over 2 EndFraction Baseline e Superscript j left-parenthesis minus k x plus omega t right-parenthesis Baseline with bold-italic k equals k left-parenthesis 1 minus j StartFraction eta Over 2 EndFraction right-parenthesis  (2.59)

This complex wavenumber naturally impacts the speed of sound

bold-italic c equals StartFraction omega Over bold-italic k EndFraction equals StartStartFraction c 0 OverOver 1 minus j StartFraction eta Over 2 EndFraction EndEndFraction  (2.60)

and the acoustic impedance

bold-italic z equals rho 0 bold-italic c equals StartStartFraction z 0 OverOver 1 minus j StartFraction eta Over 2 EndFraction EndEndFraction  (2.61)

The shown quantities of the plane wave field can also be applied in three-dimensional space and they are summarized in Table 2.2.

Table 2.2Field and energy properties of acoustic waves.

QuantitySymbolFormulaUnitsPlane waveEquation
Acoustic velocityv1jωc0pm/spρ0c0(2.35)
Acoustic impedancezp/vPa s/mz0=ρ0c0(2.38)
IntensityI12Re(pv*)Pa m/sIT=p^22ρ0c0(2.47)
Energy densityeJ/m 3eT=p^22ρ0c02(2.52)
Acoustic powerΠΠ=IAWΠT=Ap^22ρ0c0(2.43)

2.4 Fundamental Acoustic Sources

The radiation of sound is key to understanding how energy is introduced into wave fields. Depending on the wavelength, geometry, and dimension of the source the behavior varies. A detailed understanding of fundamental sources is helpful for the radiation of vibrating structures and thus, how they exchange acoustic energy.

2.4.1 Monopoles – Spherical Sources

The most simple geometry we might think of is a point in space. For simple derivation of the sound field of a point source the spherical coordinate system is introduced as shown in Figure 2.4 and defined by the following coordinate transformation

x equals r sine theta cosine phi  (2.62a)
y equals r sine theta sine phi  (2.62b)
z equals r cosine theta  (2.62c)

Figure 2.4Definition of a spherical coordinate system.
Source: Alexander Peiffer.

Using this coordinate system and neglecting the angular components the Laplace operator Δ reads as

normal upper Delta equals StartFraction 1 Over r squared EndFraction StartFraction partial-differential Over partial-differential r EndFraction left-parenthesis r squared StartFraction partial-differential Over partial-differential r EndFraction right-parenthesis equals StartFraction 2 Over r EndFraction StartFraction partial-differential Over partial-differential r EndFraction plus StartFraction partial-differential squared Over partial-differential r squared EndFraction  (2.63)

The wave equation for the velocity potential (2.30) becomes

left-parenthesis StartFraction 1 Over c 0 squared EndFraction StartFraction partial-differential squared Over partial-differential t squared EndFraction minus StartFraction 2 Over r EndFraction StartFraction partial-differential Over partial-differential r EndFraction minus StartFraction partial-differential squared Over partial-differential r squared EndFraction right-parenthesis normal upper Phi equals 0  (2.64)

The two right terms can be written in a different form using rΦ as argument

left-parenthesis StartFraction 1 Over c 0 squared EndFraction StartFraction partial-differential squared Over partial-differential t squared EndFraction minus StartFraction 1 Over r EndFraction StartFraction partial-differential squared Over partial-differential r squared EndFraction right-parenthesis left-parenthesis r normal upper Phi right-parenthesis equals 0 period  (2.65)

Equation (2.30) is the one-dimensional wave equation for the argument rΦ, so we can use the D’Alambert solution

r normal upper Phi left-parenthesis r comma t right-parenthesis equals f left-parenthesis r minus c 0 t right-parenthesis plus g left-parenthesis r plus c 0 t right-parenthesis  (2.66)

Figure 2.5Breathing sphere as source model for a monopole.
Source: Alexander Peiffer.

The first term represents an outgoing wave travelling away from the source, the second an incoming wave travelling to the source. As we are interested in sound being emitted from the source we consider the outgoing harmonic solution with complex amplitude A

normal upper Phi left-parenthesis r comma t right-parenthesis equals StartFraction bold-italic upper A Over r EndFraction e Superscript j left-parenthesis omega t minus k r right-parenthesis  (2.67)

Consider a pulsating sphere of radius R in the centre with normal surface velocity vR. With the velocity potential the radial velocity can be easily derived from the solution (2.67):

StartFraction partial-differential normal upper Phi Over partial-differential r EndFraction equals v Subscript r  (2.68)

Substituting Equation (2.67) into (2.68) and solving for A gives

bold-italic upper A equals minus bold-italic v Subscript upper R Baseline StartFraction upper R squared Over 1 minus j k upper R EndFraction e Superscript j k a  (2.69)

Hence,

normal upper Phi left-parenthesis r comma t right-parenthesis equals minus StartFraction bold-italic v Subscript upper R Baseline Over r EndFraction StartFraction upper R squared Over 1 minus j k upper R EndFraction e Superscript j left-bracket omega t minus k left-parenthesis r minus upper R right-parenthesis right-bracket  (2.70)

The strength Q(t) of the source is defined by the volume flow rate. This is the surface of the sphere times normal velocity vR

upper Q left-parenthesis t right-parenthesis equals ModifyingAbove upper V With dot equals 4 pi upper R squared v Subscript upper R Baseline left-parenthesis t right-parenthesis  (2.71)

With the harmonic source strength

StartLayout 1st Row 1st Column bold-italic upper Q left-parenthesis t right-parenthesis 2nd Column equals 4 pi upper R squared bold-italic v Subscript upper R Baseline e Superscript j omega t Baseline 3rd Column bold-italic upper Q left-parenthesis omega right-parenthesis 4th Column equals 4 pi upper R squared bold-italic v Subscript upper R EndLayout  (2.72)

the spherical wave solution is

normal upper Phi left-parenthesis r comma omega right-parenthesis equals minus StartFraction bold-italic upper Q left-parenthesis omega right-parenthesis Over 4 pi r EndFraction left-parenthesis StartFraction 1 Over 1 plus j k upper R EndFraction right-parenthesis e Superscript minus j k left-parenthesis r minus upper R right-parenthesis  (2.73)

Using equations (2.25) and (2.35) pressure and velocity are given by

bold-italic p left-parenthesis r comma omega right-parenthesis equals minus j k rho 0 c 0 normal upper Phi equals StartFraction bold-italic upper Q left-parenthesis omega right-parenthesis Over 4 pi r EndFraction left-parenthesis StartFraction j k rho 0 c 0 Over 1 plus j k upper R EndFraction right-parenthesis e Superscript minus j k left-parenthesis r minus upper R right-parenthesis  (2.74)

and

bold-italic v Subscript r Baseline left-parenthesis r comma omega right-parenthesis equals minus left-parenthesis StartFraction 1 plus j k r Over r EndFraction right-parenthesis normal upper Phi equals StartFraction bold-italic upper Q left-parenthesis omega right-parenthesis Over 4 pi r squared EndFraction left-parenthesis StartFraction 1 plus j k r Over 1 plus j k upper R EndFraction right-parenthesis e Superscript minus j k left-parenthesis r minus upper R right-parenthesis  (2.75)

2.4.1.1 Field Properties of Spherical Waves

The acoustic impedance z is according to Equation (2.38)

bold-italic z equals StartFraction j rho 0 c 0 k r Over 1 plus j k r EndFraction equals rho 0 c 0 left-brace StartFraction k squared r squared Over 1 plus k squared r squared EndFraction plus j StartFraction k r Over 1 plus k squared r squared EndFraction right-brace  (2.76)

In contrast to the plane wave, the specific acoustic impedance is not real. It contains a resistive and a reactive part. When the resistive part is dominant the pressure is in phase with the velocity. When the reactive part dominates, the velocity is out of phase to the pressure. The out of phase component does not generate any power in the sound field as it was the case for moving a mass or driving a spring. The motion is partly introduced into the local kinetic energy, and this part can be recovered as it is the case for an oscillating mass. For the acoustic field of a spherical source the reactive field represents the near-field fluid volume that is carried by the sphere motion but not emitting a wave.

Figure 2.6 Reactance and resistance of specific acoustic impedance of a pulsating sphere.
Source: Alexander Peiffer.

There are two limit cases in Equation (2.76):

  1. kr1; the wave length λ is much larger than distance r.
  2. kr1; the wave length λ is much smaller than distance r.

Introducing the above approximations into Equation (2.76) gives a fully reactive impedance for (i)

k r much-less-than 1 long right double arrow bold-italic z equals j rho 0 c 0 k r equals j rho 0 r omega  (2.77)

and a resistive part equal to plane waves for (ii)

k r much-greater-than 1 long right double arrow bold-italic z equals rho 0 c 0  (2.78)

2.4.1.2 Field Intensity, Power and Source Strength

The time averaged radiated intensity is

mathematical left-angle upper I left-parenthesis r right-parenthesis mathematical right-angle Subscript upper T Baseline equals StartFraction ModifyingAbove upper Q With caret squared k squared rho 0 c Over 32 pi squared r squared left-parenthesis 1 plus k squared upper R squared right-parenthesis EndFraction equals StartFraction upper Q Subscript r m s Superscript 2 Baseline k squared rho 0 c Over 16 pi squared r squared left-parenthesis 1 plus k squared upper R squared right-parenthesis EndFraction  (2.79)

The total radiated power can now be evaluated from Equation (2.54) and the integration surface 4πr2

mathematical left-angle normal upper Pi mathematical right-angle Subscript upper T Baseline equals 4 pi r squared mathematical left-angle upper I left-parenthesis r right-parenthesis mathematical right-angle Subscript upper T Baseline equals StartFraction upper Q Subscript r m s Superscript 2 Baseline k squared rho 0 c 0 Over 4 pi left-parenthesis 1 plus k squared upper R squared right-parenthesis EndFraction  (2.80)

The mean square pressure can be derived from (2.74) and expressed by the intensity using (2.45)

p Subscript normal r normal m normal s Superscript 2 Baseline equals one-half upper R e left-bracket bold-italic p bold-italic p Superscript asterisk Baseline right-bracket equals StartFraction upper Q Subscript r m s Superscript 2 Baseline k squared left-parenthesis rho 0 c 0 right-parenthesis squared Over 16 pi squared r squared left-parenthesis 1 plus k squared upper R squared right-parenthesis EndFraction equals mathematical left-angle upper I left-parenthesis r right-parenthesis mathematical right-angle Subscript upper T Baseline rho 0 c 0 period  (2.81)

Replacing the intensity in (2.81) gives the rms pressure in the spherical sound field due to power

p Subscript normal r normal m normal s Superscript 2 Baseline equals StartFraction rho 0 c 0 Over 4 pi r squared EndFraction mathematical left-angle normal upper Pi mathematical right-angle Subscript upper T  (2.82)

2.4.1.3 Power and Radiation Impedance at the Surface Sphere

The characteristic impedance of the sphere exactly at the surface at radius R can be translated into the radiation impedance of the sphere as a volume source. The radiation impedance is defined as the ratio of pressure to source strength at the vibrating surface

bold-italic upper Z Subscript a Baseline equals StartFraction bold-italic p Subscript normal s normal u normal r normal f Baseline Over bold-italic upper Q Subscript normal s normal u normal r normal f Baseline EndFraction equals StartFraction bold-italic p Subscript normal s normal u normal r normal f Baseline Over upper A Subscript s Baseline bold-italic v Subscript normal s normal u normal r normal f Baseline EndFraction upper A Subscript s Baseline equals 4 pi upper R squared  (2.83)

If we assume a constant harmonic surface velocity vR we get for the radiation impedance of the breathing sphere and according to the acoustic impedance (2.76)

bold-italic upper Z Subscript a Baseline equals StartFraction j rho 0 c 0 k upper R Over 4 pi upper R squared left-parenthesis 1 plus j k upper R right-parenthesis EndFraction  (2.84)

The acoustic radiation impedance is the ratio pressure and normal velocity at the sphere’s surface

bold-italic z Subscript a Baseline equals 4 pi a squared bold-italic upper Z Subscript a Baseline equals StartFraction j rho 0 c 0 k upper R Over 1 plus j k upper R EndFraction  (2.85)

We can now use this impedance to eliminate either p or vr. The power transmitted by a vibrating sphere using Equation (2.54) over the surface of the sphere

normal upper Pi equals one-half upper A Subscript s Baseline upper R e left-bracket bold-italic p bold-italic v Subscript r Superscript asterisk Baseline right-bracket equals one-half upper R e left-parenthesis bold-italic p bold-italic upper Q Superscript asterisk Baseline right-parenthesis  (2.86)
normal upper Pi equals one-half upper R e left-parenthesis bold-italic p bold-italic upper Q Superscript asterisk Baseline right-parenthesis equals one-half upper R e left-parenthesis bold-italic upper Z Subscript a Baseline bold-italic upper Q bold-italic upper Q Superscript asterisk Baseline right-parenthesis equals upper R e left-parenthesis bold-italic upper Z Subscript a Baseline right-parenthesis upper Q Subscript r m s Superscript 2  (2.87)

or for constant source pressure

normal upper Pi equals one-half upper R e left-parenthesis bold-italic p bold-italic upper Q Superscript asterisk Baseline right-parenthesis equals one-half upper R e left-parenthesis bold-italic p bold-italic p Superscript asterisk Baseline StartFraction 1 Over bold-italic upper Z Subscript a Superscript asterisk Baseline EndFraction right-parenthesis equals upper R e left-parenthesis StartFraction 1 Over bold-italic upper Z Subscript a Baseline EndFraction right-parenthesis p Subscript r m s Superscript 2  (2.88)

It is instructive to see the mechanical properties considering the limit cases from above and extract the mass that is moved by the surface. From Newtons’s law a force given by F=4πR2p leads to an in-phase acceleration of jωvr of a mass m

bold-italic upper F equals bold-italic p Baseline 4 pi upper R squared equals m j omega bold-italic v Subscript r

hence

m equals upper R e left-parenthesis StartFraction 1 Over j omega EndFraction StartFraction bold-italic p Over bold-italic upper Q EndFraction right-parenthesis equals StartFraction 1 Over omega EndFraction upper I m left-parenthesis bold-italic upper Z Subscript a Baseline right-parenthesis  (2.89)

For kR1 we get: m=4πR3ρ0=3Vsphρ0. Thus, at low frequencies the source surface motion carries three times the fluid volume of the sphere. This motion near the source is called an evanescent wave, because it is oscillatory motion of fluid that does not radiate.

2.4.1.4 Point Sources

A point source is a spherical source with an infinitely small radius. Performing the limit kR0 for Equation (2.73) leads to the velocity potential for point sources of strength Q

normal upper Phi left-parenthesis r comma omega right-parenthesis equals minus StartFraction bold-italic upper Q left-parenthesis omega right-parenthesis Over 4 pi r EndFraction e Superscript minus j k r  (2.90)

The pressure and velocity field of such a source is given by

bold-italic p left-parenthesis r comma omega right-parenthesis equals minus j k rho 0 c 0 normal upper Phi equals StartFraction j k rho 0 c 0 bold-italic upper Q left-parenthesis omega right-parenthesis Over 4 pi r EndFraction e Superscript minus j k r  (2.91)

and

bold-italic v Subscript r Baseline left-parenthesis r comma omega right-parenthesis equals StartFraction partial-differential normal upper Phi Over partial-differential r EndFraction normal upper Phi equals StartFraction bold-italic upper Q left-parenthesis omega right-parenthesis Over 4 pi r EndFraction j k left-parenthesis 1 plus StartFraction 1 Over j k r EndFraction right-parenthesis e Superscript minus j k r Baseline equals Overscript r much-greater-than k Endscripts StartFraction j k bold-italic upper Q Over 4 pi r EndFraction e Superscript minus j k r  (2.92)

All other relations regarding power and intensity expressions remain. We see that the limit is expressed for kR and not for the wavelength. The reason is that it is the ratio of a characteristic length (in this case the sphere radius) to the wavelength that determines if the geometrical details must be considered or not. In other words, a wave of a certain wavelength doesn’t care about details that are much smaller.

With the D’Alambert solution for spherical waves (2.66) we can also derive a point source in time domain

normal upper Psi left-parenthesis r comma t right-parenthesis equals minus StartFraction upper Q left-parenthesis t minus c 0 slash r right-parenthesis Over 4 pi r EndFraction  (2.93)

The point source is of great importance for the solution of the inhomogeneous wave equation in combination with complex boundary conditions. Any source can be reconstructed by a superposition of point sources as shown in Section 2.7.

Performing the limit process with kR0 and taking the power from equation 2.86 we get the intensity of the point source:

mathematical left-angle upper I left-parenthesis r right-parenthesis mathematical right-angle Subscript upper T Baseline equals StartFraction upper Q Subscript rms Superscript 2 Baseline k squared rho 0 c Over 16 pi squared r squared EndFraction  (2.94)

and the total radiated power

mathematical left-angle normal upper Pi mathematical right-angle Subscript upper T Baseline equals StartFraction upper Q Subscript rms Superscript 2 Baseline k squared rho 0 c 0 Over 4 pi EndFraction  (2.95)

with radiation impedance following from this

upper Z Subscript a Baseline equals StartFraction k squared rho 0 c 0 Over 4 pi EndFraction  (2.96)

2.5 Reflection of Plane Waves

A plane wave striking a plane surface is a first example of interaction with obstacles. Imagine a configuration as shown in Figure 2.7. The impedance of the surface is z2, and it is given by using the velocity vz perpendicular to the plane.

Figure 2.7 Reflection of a plane wave at an infinite surface with impedance Z2.
Source: Alexander Peiffer.

Without loss of generality the wave front is parallel to the y-axis and all properties are functions of x and z. The solution in the half space of z>0 is the superposition of two plane waves.

normal upper Phi 1 left-parenthesis r right-parenthesis equals normal upper Phi e Superscript minus j bold k bold r Baseline plus normal upper Phi Superscript left-parenthesis upper R right-parenthesis Baseline e Superscript minus j bold k Super Superscript left-parenthesis upper R right-parenthesis Superscript bold r  (2.97)

With the following arguments of the exponential function

StartLayout 1st Row 1st Column bold k bold r 2nd Column equals k left-parenthesis sine theta x minus cosine theta z right-parenthesis EndLayout  (2.98)
StartLayout 1st Row 1st Column bold k Superscript left-parenthesis upper R right-parenthesis Baseline bold r 2nd Column equals k left-parenthesis sine theta Superscript left-parenthesis upper R right-parenthesis Baseline x plus cosine theta Superscript left-parenthesis upper R right-parenthesis Baseline z right-parenthesis EndLayout  (2.99)

The pressure at the surface z=0 is given by

bold-italic p left-parenthesis x comma z equals 0 right-parenthesis equals j omega rho 0 normal upper Phi 1 equals j omega rho 0 left-parenthesis normal upper Phi e Superscript minus j k sine theta x Baseline plus normal upper Phi Superscript left-parenthesis upper R right-parenthesis Baseline e Superscript minus j k sine theta Super Superscript left-parenthesis upper R right-parenthesis Superscript x Baseline right-parenthesis  (2.100)

and the velocity in z-direction reads

StartLayout 1st Row 1st Column bold-italic v Subscript z Baseline left-parenthesis x comma z equals 0 right-parenthesis 2nd Column equals minus StartFraction partial-differential normal upper Phi Over partial-differential z EndFraction 2nd Row 1st Column Blank 2nd Column equals j k left-parenthesis cosine theta normal upper Phi e Superscript minus j k sine theta x Baseline minus cosine theta Superscript left-parenthesis upper R right-parenthesis Baseline normal upper Phi Superscript left-parenthesis upper R right-parenthesis Baseline e Superscript minus j k sine theta Super Superscript left-parenthesis upper R right-parenthesis Superscript x Baseline right-parenthesis EndLayout  (2.101)

We certainly shall not be able to match the impedance z2=p/vz at every surface position unless the arguments of the exponential functions are equal, hence

theta equals theta Superscript left-parenthesis upper R right-parenthesis

So, we get from the surface impedance condition

bold-italic z 2 left-parenthesis x comma z equals 0 right-parenthesis equals StartFraction bold-italic p left-parenthesis x comma z equals 0 right-parenthesis Over bold-italic v Subscript z Baseline left-parenthesis x comma z equals 0 right-parenthesis EndFraction equals StartFraction z 0 left-parenthesis normal upper Phi plus normal upper Phi Superscript left-parenthesis upper R right-parenthesis Baseline right-parenthesis Over cosine theta left-parenthesis normal upper Phi minus normal upper Phi Superscript left-parenthesis upper R right-parenthesis Baseline right-parenthesis EndFraction  (2.102)

With z0=ρ0c0 and rearranging the above equation, the reflection factor is given by

bold-italic upper R equals StartFraction normal upper Phi Superscript left-parenthesis upper R right-parenthesis Baseline Over normal upper Phi EndFraction equals StartFraction bold-italic z 2 cosine theta minus z 0 Over bold-italic z 2 cosine theta plus z 0 EndFraction  (2.103)

The ratio between irradiated power to reflector power is the squared reflection factor called the.

StartLayout 1st Row 1st Column r Subscript s Baseline left-parenthesis theta right-parenthesis 2nd Column upper R squared left-parenthesis theta right-parenthesis equals StartFraction left-parenthesis bold-italic z 2 cosine theta minus z 0 right-parenthesis squared Over left-parenthesis bold-italic z 2 cosine theta plus z 0 right-parenthesis squared EndFraction EndLayout  (2.104)
StartLayout 1st Row 1st Column alpha Subscript normal s Baseline left-parenthesis theta right-parenthesis 2nd Column 1 minus r Subscript s Baseline left-parenthesis theta right-parenthesis equals left-parenthesis 1 minus StartAbsoluteValue upper R left-parenthesis theta right-parenthesis EndAbsoluteValue squared right-parenthesis EndLayout  (2.105)

Note that those coefficients are exclusively described by the impedance of fluid and surface and not density or speed of sound. Thus, the impedance is the relevant quantity here.

2.6 Reflection and Transmission of Plane Waves

A plane wave passing a flat interface between two infinite fluid volumes with different density and sound velocity as shown in Figure 2.8 is a first example of continuous systems exchanging acoustic energy. Applications of such a system could be for example the interface between a liquid (water) and a gas (air) or just different gases.

Figure 2.8Transmission and reflection of a plane wave at the interface of two fluids.
Source: Alexander Peiffer.

Region 1 of the incoming wave has two wave components, the incoming and the reflected wave, and region 2 the transmitted wave. Thus, both velocity potentials read

StartLayout 1st Row 1st Column normal upper Phi 1 left-parenthesis bold r 1 right-parenthesis 2nd Column equals normal upper Phi 1 e Superscript minus j bold k 1 bold r 1 Baseline plus normal upper Phi 1 Superscript left-parenthesis upper R right-parenthesis Baseline e Superscript minus j bold k 1 Super Superscript left-parenthesis upper R right-parenthesis Superscript bold r 1 Baseline 2nd Row 1st Column normal upper Phi 2 left-parenthesis bold r 2 right-parenthesis 2nd Column equals normal upper Phi 2 e Superscript minus j bold k 2 bold r 2 EndLayout  (2.106)

Using the given angles as sketched in Figure 2.8 the wavenumber space vector products are given by

StartLayout 1st Row 1st Column bold k 1 bold r 1 2nd Column equals k 1 left-parenthesis sine theta 1 x minus cosine theta 1 z right-parenthesis EndLayout  (2.107)
StartLayout 1st Row 1st Column bold k 1 bold r 1 Superscript left-parenthesis upper R right-parenthesis 2nd Column equals k 1 left-parenthesis sine theta 1 Superscript left-parenthesis upper R right-parenthesis Baseline x plus cosine theta 1 Superscript left-parenthesis upper R right-parenthesis Baseline z right-parenthesis EndLayout  (2.108)
StartLayout 1st Row 1st Column bold k 2 bold r 2 2nd Column k 2 left-parenthesis sine theta 2 x minus cosine theta 2 z right-parenthesis equals l a b e l less-than 2.107 greater-than slash l a b e l less-than EndLayout  (2.109)

with k1=ω/c1 and k2=ω/c2. The contact face between the fluid requires the continuity of pressure and velocity in the z-direction. We start with the pressure p1/2=jωρ1/2Φ1/2. Entering equations (2.107)–(2.109) into (2.106) and determining the pressure relation

p 1 left-parenthesis x comma z equals 0 right-parenthesis equals p 2 left-parenthesis x comma z equals 0 right-parenthesis

gives

rho 1 normal upper Phi 1 e Superscript j k 1 sine theta 1 x Baseline plus rho 1 normal upper Phi 1 Superscript left-parenthesis upper R right-parenthesis Baseline e Superscript j k 1 sine theta 1 Super Superscript left-parenthesis upper R right-parenthesis Superscript x Baseline equals rho 2 normal upper Phi 2 e Superscript j k 2 sine theta 2 x  (2.110)

A solution for any x is only possible if the arguments of the exponential functions are equal.

k 1 sine theta 1 equals k 1 sine theta 1 Superscript left-parenthesis upper R right-parenthesis  (2.111)

So also in the transmission case the incident angle equals the angle of the reflected wave. Additionally we have

k 1 sine theta 1 equals k 2 sine theta 2  (2.112)

This represents the acoustic equivalent of Snell’s law of transmission:

StartFraction sine theta 1 Over c 1 EndFraction equals StartFraction sine theta 2 Over c 2 EndFraction  (2.113)

With these conditions we can factor out the exponential function

rho 1 left-parenthesis normal upper Phi 1 plus normal upper Phi 1 Superscript left-parenthesis upper R right-parenthesis Baseline right-parenthesis equals rho 2 normal upper Phi 2  (2.114)

After clarifying the angles of reflection and transmission the next point is to assess the fraction of transmitted and reflected wave. From the continuity condition for the velocity in the z-direction we get with vz=Φ/z when going through the algebra

StartFraction sine theta 1 Over c 1 EndFraction left-parenthesis normal upper Phi 1 minus normal upper Phi 1 Superscript left-parenthesis upper R right-parenthesis Baseline right-parenthesis equals StartFraction sine theta 2 Over c 2 EndFraction normal upper Phi 2  (2.115)

Rearranging equations (2.114) and (2.115) the reflection factor is

bold-italic upper R equals StartFraction normal upper Phi 1 Superscript left-parenthesis upper R right-parenthesis Baseline Over normal upper Phi 1 EndFraction equals StartStartFraction StartFraction z 2 Over cosine theta 2 EndFraction minus StartFraction z 1 Over cosine theta 1 EndFraction OverOver StartFraction z 2 Over cosine theta 2 EndFraction plus StartFraction z 1 Over cosine theta 1 EndFraction EndEndFraction  (2.116)

This expression is similar to (2.103) except the angle factor that represents the physics of the wave propagation in the second medium. The transmission factor is defined by the ratio of pressure amplitudes p1=jωρ1Φ1 and p2=jωρ2Φ2. Hence,

bold-italic upper T equals StartStartFraction StartFraction 2 z 2 Over cosine theta 2 EndFraction OverOver StartFraction z 1 Over cosine theta 1 EndFraction plus StartFraction z 2 Over cosine theta 2 EndFraction EndEndFraction  (2.117)

The transmitted acoustic power follows from the square of the amplitudes. We introduce a transmission coefficient τ by

tau equals StartFraction normal upper Pi Subscript trans Baseline Over normal upper Pi Subscript in Baseline EndFraction  (2.118)

Without loss of generality we assume ϑ=0 so each power is given by

StartLayout 1st Row 1st Column normal upper Pi Subscript in 2nd Column equals StartFraction upper A Over 2 EndFraction upper R e left-parenthesis StartFraction ModifyingAbove p With caret Subscript 1 Superscript 2 Baseline Over bold-italic z 1 EndFraction right-parenthesis 3rd Column normal upper Pi Subscript trans 4th Column StartFraction upper A Over 2 EndFraction upper R e left-parenthesis StartFraction ModifyingAbove p With caret Subscript 2 Superscript 2 Baseline Over bold-italic z 2 EndFraction right-parenthesis equals StartFraction upper A Over 2 EndFraction upper R e left-parenthesis StartFraction bold-italic upper T squared ModifyingAbove p With caret Subscript 1 Superscript 2 Baseline Over bold-italic z 2 EndFraction right-parenthesis EndLayout

With (2.117) this reads as:

tau equals StartFraction 4 upper R e left-parenthesis bold-italic z 1 right-parenthesis upper R e left-parenthesis bold-italic z 2 right-parenthesis Over StartAbsoluteValue bold-italic z 1 plus bold-italic z 2 EndAbsoluteValue squared EndFraction  (2.120)

It should be noted that the transmission coefficient of the flat interface between two fluids is determined by the impedance of each half space ‘seen’ from the other side. This is the first indication for the coupling of subsystems determined by the radiation impedance into the free fields of each subsystem. Similar expressions will be found in Sec. 8.2.4.1 when transmission is dealt with in the context of coupled random subsystems.

2.7 Inhomogeneous Wave Equation

In the considerations in this chapter so far, we neglected the source terms related to the conservation of mass and momentum. All sources discussed until now are caused by vibrating surfaces. For establishing a physical link between the source term and the specific mass flow q˙s in Equation (2.3) and force density term f in Equation (2.8) we keep the terms this time. The source terms are not influenced by the linearization procedure; thus, the inhomogeneous and linear equations of momentum (2.24) and continuity (2.23) read as

StartLayout 1st Row 1st Column StartFraction partial-differential rho prime Over partial-differential t EndFraction plus rho 0 nabla bold v prime 2nd Column equals ModifyingAbove rho With dot Subscript normal s Baseline 3rd Column rho 0 StartFraction partial-differential bold v prime Over partial-differential t EndFraction plus nabla p 4th Column equals bold f EndLayout  (2.121)

Repeating the steps of section 2.2.5 we finally get the inhomogeneous wave equation

StartFraction 1 Over c 0 squared EndFraction StartFraction partial-differential squared p Over partial-differential t squared EndFraction minus nabla squared p equals ModifyingAbove rho With two-dots Subscript s Baseline minus nabla bold f  (2.122)

The density source is converted into a volume source strength density by ρ˙s=ρ0qs(t). The above equation can also be converted into the frequency domain and hence to the inhomogeneous Helmholtz equation

left-parenthesis k squared plus normal upper Delta right-parenthesis bold-italic p left-parenthesis bold x comma omega right-parenthesis equals minus j omega rho 0 bold-italic q Subscript s Baseline plus nabla bold f equals minus bold-italic f Subscript q Baseline left-parenthesis bold r right-parenthesis  (2.123)

2.7.1 Acoustic Green’s Functions

This section presents the concept of the Green’s function that uses a formalism to calculate the sound field for arbitrary source and boundary configurations as shown for example by Morse and Ingard (1968).

The Green’s function is defined as the solution of the following inhomogeneous wave equation.

normal upper Delta g left-parenthesis bold r comma bold r 0 right-parenthesis plus k squared g left-parenthesis bold r comma bold r 0 right-parenthesis equals minus delta left-parenthesis bold r minus bold r 0 right-parenthesis  (2.124)

The inhomogeneous part is the delta function which allows for this elegant derivation of the Kirchhoff integral. The delta function is introduced in the appendix A.1.3 in the time domain. However, it can also be applied in space. The multidimensional delta function is simply the product of three Dirac delta functions in space

delta left-parenthesis bold r minus bold r 0 right-parenthesis equals delta left-parenthesis x minus x 0 right-parenthesis delta left-parenthesis y minus y 0 right-parenthesis delta left-parenthesis y minus y 0 right-parenthesis with bold r equals Start 1 By 1 Matrix 1st Row x comma y comma z EndMatrix Superscript upper T  (2.125)

The sifting properties and the value of the integration is defined by volume integral

f left-parenthesis bold r right-parenthesis equals integral Underscript upper V Endscripts f left-parenthesis bold r 0 right-parenthesis delta left-parenthesis bold r minus bold r 0 right-parenthesis d bold r 0 and integral Underscript upper V Endscripts delta left-parenthesis bold r right-parenthesis d bold r equals 1  (2.126)

The solution of Equation (2.124) is the point source (2.91)3.

g left-parenthesis bold r vertical-bar bold r 0 right-parenthesis equals StartFraction 1 Over 4 pi l EndFraction e Superscript minus j k l Baseline with l equals StartAbsoluteValue bold r minus bold r 0 EndAbsoluteValue  (2.127)

In order to achieve a common formulation we add an arbitrary solution χ of the homogeneous wave equation

normal upper Delta chi left-parenthesis bold r vertical-bar bold r 0 right-parenthesis plus k squared chi left-parenthesis bold r vertical-bar bold r 0 right-parenthesis equals 0  (2.128)

to the Green’s function to get the generalized Green’s function

upper G left-parenthesis bold r comma bold r 0 right-parenthesis equals g left-parenthesis bold r comma bold r 0 right-parenthesis plus chi left-parenthesis bold r comma bold r 0 right-parenthesis  (2.129)

The purpose of the additional homogeneous solution is to create freedom to fulfill boundary conditions that do not occur in the free sound field. The task is to find the solution for the inhomogeneous wave equation

normal upper Delta bold-italic p left-parenthesis bold r right-parenthesis plus k squared bold-italic p left-parenthesis bold r right-parenthesis equals minus bold-italic f Subscript q Baseline left-parenthesis bold r right-parenthesis  (2.130)

The generalized Green’s function must be a solution of the following equation for the special case with r,r0V and the boundary V as shown in Figure 2.9.

normal upper Delta upper G left-parenthesis bold r comma bold r 0 right-parenthesis plus k squared upper G left-parenthesis bold r comma bold r 0 right-parenthesis equals minus delta left-parenthesis bold r minus bold r 0 right-parenthesis  (2.131)

Figure 2.9 Solution volume and boundaries.
Source: Alexander Peiffer.

In order to receive a global solution we perform the operation

upper G left-parenthesis bold r comma bold r 0 right-parenthesis dot left-parenthesis 2.130 right-parenthesis minus bold-italic p left-parenthesis bold r right-parenthesis dot left-parenthesis 2.131 right-parenthesis  (2.132)

This leads to

upper G left-parenthesis bold r comma bold r 0 right-parenthesis normal upper Delta bold-italic p left-parenthesis bold r right-parenthesis minus bold-italic p left-parenthesis bold r right-parenthesis normal upper Delta upper G left-parenthesis bold r comma bold r 0 right-parenthesis equals minus left-bracket upper G left-parenthesis bold r comma bold r 0 right-parenthesis bold-italic f Subscript q Baseline left-parenthesis bold r right-parenthesis minus bold-italic p left-parenthesis bold r right-parenthesis delta left-parenthesis bold r comma bold r 0 right-parenthesis right-bracket  (2.133)

Exchanging r and r0 and integrating r0 over the volume V gives

StartLayout 1st Row 1st Column integral Underscript upper V Endscripts upper G left-parenthesis bold r 0 comma bold r right-parenthesis normal upper Delta bold-italic p left-parenthesis bold r 0 right-parenthesis 2nd Column bold-italic p left-parenthesis bold r 0 right-parenthesis normal upper Delta upper G left-parenthesis bold r 0 comma bold r right-parenthesis d bold r 0 equals 2nd Row 1st Column Blank 2nd Column integral Underscript upper V Endscripts upper G left-parenthesis bold r 0 comma bold r right-parenthesis bold-italic f Subscript q Baseline left-parenthesis bold r 0 right-parenthesis d bold r 0 plus ModifyingBelow integral Underscript upper V Endscripts bold-italic p left-parenthesis bold r 0 right-parenthesis delta left-parenthesis bold r minus bold r 0 right-parenthesis d bold r 0 With bottom-brace Underscript equals bold-italic p left-parenthesis bold r right-parenthesis Endscripts EndLayout  (2.134)

The last term on the RHS follows from the sifting property of the delta function

StartLayout 1st Row 1st Column bold-italic p left-parenthesis bold r right-parenthesis 2nd Column equals integral Underscript upper V Endscripts upper G left-parenthesis bold r 0 comma bold r right-parenthesis bold-italic f Subscript q Baseline left-parenthesis bold r right-parenthesis d bold r 0 2nd Row 1st Column Blank 2nd Column integral Underscript upper V Endscripts upper G left-parenthesis bold r 0 comma bold r right-parenthesis normal upper Delta bold-italic p left-parenthesis bold r 0 right-parenthesis minus bold-italic p left-parenthesis bold r 0 right-parenthesis normal upper Delta upper G left-parenthesis bold r 0 comma bold r right-parenthesis d bold r 0 period EndLayout  (2.135)

With Green’s law of vector analysis

integral Underscript upper V Endscripts left-parenthesis normal upper Phi normal upper Delta normal upper Psi minus normal upper Psi normal upper Delta normal upper Phi right-parenthesis d upper V equals integral Underscript partial-differential upper V Endscripts left-parenthesis normal upper Phi nabla normal upper Psi minus normal upper Psi nabla normal upper Phi right-parenthesis d upper S  (2.136)

some volume integrals can be transferred into surface integrals and we get finally

StartLayout 1st Row 1st Column bold-italic p left-parenthesis bold r right-parenthesis 2nd Column equals integral Underscript upper V Endscripts upper G left-parenthesis bold r 0 comma bold r right-parenthesis bold-italic f Subscript q Baseline left-parenthesis bold r 0 right-parenthesis d bold r 0 2nd Row 1st Column Blank 2nd Column integral Underscript partial-differential upper V Endscripts upper G left-parenthesis bold r 0 comma bold r right-parenthesis nabla bold-italic p left-parenthesis bold r 0 right-parenthesis minus bold-italic p left-parenthesis bold r 0 right-parenthesis nabla upper G left-parenthesis bold r 0 comma bold r right-parenthesis d bold r 0 EndLayout  (2.137)

The first term on the right-hand side is the volume integral over all sources fq(r) in the volume. So given a known source distribution we can calculate the according sound field. The two terms in the surface integral take care of the boundary condition. The pressure gradient in the first can be converted into the normal velocity using (2.35). The second surface integral allows establishing the correct surface impedance. Equation (2.137) is called the constant frequency version of the .

2.7.2 Rayleigh integral

The Rayleigh integral is a special solution of the Kirchhoff-Helmholtz integral applied to flat and infinite surfaces. We assume a configuration as shown in Figure 2.10. The integration volume is the right half space for z>0 closed by a half sphere of infinite radius. The Green’s function of any source at r0=(x0,y0,z0) with z0>0 is as defined in equation (2.127). The rigid surface acts as a reflector. Thus, there is a mirror source located at r0=(x0,y0,z0). This source is not in volume V, and the added wave field is therefore considered as a homogeneous solution in the volume. Hence, we get for the generalized Green’s function

upper G left-parenthesis bold r comma bold r 0 right-parenthesis equals StartFraction 1 Over 4 pi l EndFraction e Superscript minus j k l Baseline plus StartFraction 1 Over 4 pi l Superscript prime Baseline EndFraction e Superscript minus j k l Super Superscript prime Superscript Baseline with l equals StartAbsoluteValue bold r minus bold r 0 EndAbsoluteValue l Superscript prime Baseline equals StartAbsoluteValue bold r minus bold r 0 prime EndAbsoluteValue  (2.138)

Figure 2.10 Half space in front of a rigid wall.
Source: Alexander Peiffer.

We enter this version of the Green’s function in Equation (2.137) and we get

StartLayout 1st Row 1st Column bold-italic p left-parenthesis bold r right-parenthesis 2nd Column equals integral Underscript upper V Endscripts upper G left-parenthesis bold r 0 comma bold r right-parenthesis f Subscript q Baseline left-parenthesis bold r 0 right-parenthesis d bold r 0 2nd Row 1st Column Blank 2nd Column integral Subscript negative normal infinity Superscript normal infinity Baseline integral Subscript negative normal infinity Superscript normal infinity Baseline upper G left-parenthesis bold r 0 comma bold r right-parenthesis StartFraction partial-differential bold-italic p left-parenthesis bold r 0 right-parenthesis Over partial-differential z EndFraction minus bold-italic p left-parenthesis bold r 0 right-parenthesis StartFraction partial-differential upper G left-parenthesis bold r 0 comma bold r right-parenthesis Over partial-differential z EndFraction d x 0 d y 0 period EndLayout  (2.139)

We assume a source-free half space so fq(r)=0, and due to the mirror source symmetry G(r0,r)z=0 is also true. By clever selection of the Green’s function we fulfilled the boundary condition automatically. For the surface integral the contributions from the half sphere with infinite radius are supposed to be zero. From Equation (2.35) the first expression can be converted into an expression for the surface velocity vz. Performing the limit process z00 we get

upper G left-parenthesis bold r 0 comma bold r right-parenthesis equals StartFraction 2 Over 4 pi l EndFraction e Superscript minus j k l Baseline l equals StartRoot z squared plus left-parenthesis x minus x 0 right-parenthesis squared plus left-parenthesis y minus y 0 right-parenthesis squared EndRoot  (2.140)

and with this Green’s function we can derive the Rayleigh integral that allows the calculation of infinite half space sound fields excited by a rigid vibrating plane with arbitrary velocity distribution vz(x0,y0).

bold-italic p left-parenthesis bold r right-parenthesis equals integral Subscript negative normal infinity Superscript normal infinity Baseline integral Subscript negative normal infinity Superscript normal infinity Baseline StartFraction j omega rho 0 Over 2 pi l EndFraction e Superscript minus j k l Baseline bold-italic v Subscript z Baseline left-parenthesis x 0 comma y 0 right-parenthesis d x 0 d y 0  (2.141)

2.7.3 Piston in a Wall

A cylindrical loudspeaker in a wall can be modelled by a piston of radius R vibrating with velocity vz located in a rigid wall. For convenience the surface integral will be expressed in cylindrical coordinates r0 and φ0. The receiver coordinates are given as spherical coordinates r and ϑ(Figure 2.11). Without loss of generality the azimuthal angle φ is set to zero.

StartLayout 1st Row 1st Column bold-italic p left-parenthesis bold r comma theta right-parenthesis 2nd Column equals integral Subscript 0 Superscript 2 pi Baseline integral Subscript 0 Superscript normal infinity Baseline StartFraction j omega rho 0 Over 2 pi l EndFraction e Superscript minus j k l Baseline bold-italic v Subscript z Baseline left-parenthesis r 0 right-parenthesis r 0 d r 0 d phi 0 2nd Row 1st Column Blank 2nd Column equals integral Subscript 0 Superscript 2 pi Baseline integral Subscript 0 Superscript upper R Baseline StartFraction j omega rho 0 Over 2 pi l EndFraction e Superscript minus j k l Baseline bold-italic v Subscript z Baseline r 0 d r 0 d phi 0 EndLayout  (2.142)

Figure 2.11Coordinate definitions for the piston in the wall.
Source: Alexander Peiffer.

In the far field approximation we assume lr and get

l equals r plus r 0 sine theta cosine phi 0 period  (2.143)

So, the approximate result is

bold-italic p left-parenthesis r comma theta right-parenthesis equals StartFraction j omega rho 0 Over 2 pi r bold-italic v Subscript z Baseline EndFraction integral Subscript 0 Superscript 2 pi Baseline integral Subscript 0 Superscript upper R Baseline e Superscript j k r 0 sine theta cosine phi 0 Baseline d phi 0 r 0 d r 0  (2.144)

The integral is the Bessel function of first order

bold-italic p left-parenthesis r comma theta right-parenthesis equals StartFraction j omega rho 0 Over 2 pi r bold-italic v Subscript z Baseline EndFraction left-parenthesis StartFraction 2 upper J 1 left-parenthesis k upper R sine theta right-parenthesis Over k upper R sine theta EndFraction right-parenthesis period  (2.145)

The results for some values of kR are shown in Figure 2.12 over the angular range of ±π/2. For a piston size small compared to the wavelength (kR1) the radiation pattern is similar to a point source. The smaller the wavelength gets in relation to the piston radius R the more a specific radiation pattern develops.

Figure 2.12Angular distribution (radiation pattern) of the pressure field of the piston.
Source: Alexander Peiffer.

2.7.3.1 Impedance Concept

The radiation impedance of the piston is calculated from the pressure averaged over the surface related to the piston velocity vz. As shown by Lerch and Landes (2012) the mechanical impedance of the piston due to radiation is given by

bold-italic upper Z equals StartFraction 1 Over bold-italic v Subscript z Baseline EndFraction integral Underscript upper S Endscripts p left-parenthesis r right-parenthesis d upper S equals StartFraction 2 pi Over bold-italic v Subscript z Baseline EndFraction integral Subscript 0 Superscript upper R Baseline bold-italic p left-parenthesis r right-parenthesis r d r  (2.146)

According to equation (2.141) assuming a constant velocity vz over the surface A the pressure is

bold-italic p left-parenthesis r right-parenthesis equals StartFraction j omega rho 0 bold-italic v Subscript z Baseline Over 2 pi EndFraction integral Underscript upper A Endscripts StartFraction e Superscript minus j k s Baseline Over s EndFraction d upper A period  (2.147)

Thus, we get the pressure at r from integrating the contribution from the rest of the piston in circles of radius s. The angle integration over φ0 runs from 0 to 2π. From every angle φ0 follows the integration limits smax of the second integral.

s Subscript m a x Baseline equals r cosine phi plus StartRoot upper R squared minus r squared sine squared phi EndRoot  (2.148)

Using those limits gives

StartLayout 1st Row 1st Column bold-italic p left-parenthesis r right-parenthesis 2nd Column equals StartFraction j omega rho Subscript o Baseline bold-italic v Subscript z Baseline Over 2 pi EndFraction integral Subscript 0 Superscript 2 pi Baseline integral Subscript 0 Superscript r cosine phi plus StartRoot upper R squared minus r squared sine squared phi EndRoot Baseline StartFraction e Superscript minus j k s Baseline Over s EndFraction s d s d phi 2nd Row 1st Column Blank 2nd Column equals StartFraction rho 0 c 0 bold-italic v Subscript z Baseline Over 2 pi EndFraction integral Subscript 0 Superscript 2 pi Baseline left-parenthesis 1 minus e Superscript minus j k r cosine phi minus j k StartRoot upper R squared minus r squared sine squared phi EndRoot Baseline right-parenthesis d phi 3rd Row 1st Column Blank 2nd Column equals rho 0 c 0 bold-italic v Subscript z Baseline left-parenthesis 1 minus StartFraction 1 Over 2 pi EndFraction integral Subscript 0 Superscript 2 pi Baseline e Superscript minus j k r cosine phi minus j k StartRoot upper R squared minus r squared sine squared phi EndRoot Baseline d phi right-parenthesis period EndLayout  (2.149)

Inserting equation (2.149) into (2.146) leads to the expression

bold-italic upper Z equals rho 0 c 0 pi upper R squared left-parenthesis 1 minus StartFraction 1 Over pi upper R squared EndFraction integral Subscript phi equals 0 Superscript 2 pi Baseline integral Subscript r equals 0 Superscript upper R Baseline e Superscript minus j k r cosine phi minus j k StartRoot upper R squared minus r squared sine squared phi EndRoot Baseline r d r d phi right-parenthesis period  (2.150)

Figure 2.13Surface integration over piston for radiation impedance.
Source: Alexander Peiffer.

Running through quite a lot of algebraic modifications we get the expression for the impedance of a piston

bold-italic upper Z equals rho 0 c 0 pi upper R squared left-parenthesis 1 minus StartFraction upper J 1 left-parenthesis 2 k upper R right-parenthesis Over k upper R EndFraction plus j StartFraction upper H 1 left-parenthesis 2 k upper R right-parenthesis Over k upper R EndFraction right-parenthesis  (2.151)

or

bold-italic z Subscript a Baseline equals rho 0 c 0 left-parenthesis 1 minus StartFraction upper J 1 left-parenthesis 2 k upper R right-parenthesis Over k upper R EndFraction plus j StartFraction upper H 1 left-parenthesis 2 k upper R right-parenthesis Over k upper R EndFraction right-parenthesis  (2.152)

H1(z) is the Hankel function of first order. In Figure 2.14 the real and imaginary parts of the acoustic radiation impedance are compared to those of the pulsating sphere. Both sources have a similar shape except some waviness for the piston resulting from interference effects from the integration over the piston surface. For large kR the impedance is real for both radiators and approaches the acoustic impedance of a plane wave z0=ρ0c0.

Figure 2.14 Acoustic radiation impedance of the piston.
Source: Alexander Peiffer.

With Equation (2.87) the radiated power of a piston of source strength Q=πR2vz is

mathematical left-angle normal upper Pi mathematical right-angle Subscript upper T Baseline equals rho 0 c 0 left-parenthesis 1 minus StartFraction upper J 1 left-parenthesis 2 k upper R right-parenthesis Over k upper R EndFraction plus j StartFraction upper H 1 left-parenthesis 2 k upper R right-parenthesis Over k upper R EndFraction right-parenthesis upper Q Subscript rms Superscript 2  (2.153)

The main use of Equation (2.153) is that the required velocity to achieve (or prevent) a certain sound power can be calculated from it, for example if one must define the boundary condition for a radiating piston in simulation software and only the radiated power is known.

2.7.3.2 Inertia Effects

The Bessel functions can be approximated by a series in 2kR taking the first series term of both functions (Jacobsen, 2011)

bold-italic upper Z equals rho 0 c 0 pi upper R squared left-parenthesis one-half left-parenthesis 2 k upper R right-parenthesis squared plus j StartFraction 8 Over 3 pi EndFraction k upper R right-parenthesis  (2.154)

This expression is valid for ka<0.5. From the imaginary part we get for the mass

m equals StartFraction upper I m left-parenthesis bold-italic upper Z right-parenthesis Over omega EndFraction equals StartFraction 8 upper R cubed rho 0 Over 3 EndFraction  (2.155)

Assuming a cylindrical volume V=πR2lc of the fluid above the piston we can calculate the length of the moving mass cylinder to be

l Subscript c Baseline equals StartFraction 8 upper R Over 3 pi EndFraction almost-equals 0.85 upper R  (2.156)

meaning that at low frequencies the piston is moving a fluid layer of 0.85 times the radius acting as an inertia without radiation.

2.7.4 Power Radiation

For the radiated power calculation of the piston we took the pressure at the piston surface and integrated the pressure–velocity product over the surface. Due to the fact that the velocity is constant the surface integral involves mainly the pressure as a space-dependent property. In case of vibrating structures with complex shapes of vibration the velocity distribution over the surface is not homogeneous, and we need a more detailed approach.

bold-italic p left-parenthesis bold r right-parenthesis equals double-integral Subscript negative normal infinity Superscript normal infinity Baseline StartFraction j omega rho 0 Over 2 pi l EndFraction e Superscript minus j k l Baseline bold-italic v Subscript z Baseline left-parenthesis bold r 0 right-parenthesis d bold r 0 with l equals StartAbsoluteValue bold r minus bold r 0 EndAbsoluteValue  (2.157)

In the above equation a function with argument (rr0) is multiplied by the velocity function for r0 and integrated over the two-dimensional space. Mathematically, this can be interpreted as a two-dimensional convolution in space

bold-italic p left-parenthesis bold r right-parenthesis equals StartFraction j omega rho 0 Over 2 pi StartAbsoluteValue bold r EndAbsoluteValue EndFraction e Superscript j k StartAbsoluteValue bold r EndAbsoluteValue Baseline asterisk bold-italic v Subscript z Baseline left-parenthesis bold r right-parenthesis  (2.158)

Thus, when we apply the two-dimensional Fourier transform to the Rayleigh integral the result is the product of the Fourier transform of the vibration shape vz(r0) and the Green’s function in wavenumber space leading to

bold-italic p left-parenthesis bold k right-parenthesis equals StartFraction 1 Over 4 pi EndFraction StartFraction rho 0 omega Over StartRoot k Subscript a Superscript 2 Baseline minus k squared EndRoot EndFraction bold-italic v Subscript z Baseline left-parenthesis bold k right-parenthesis with k equals StartAbsoluteValue bold k EndAbsoluteValue  (2.159)

So, we have replaced the expensive convolution operation by a multiplication. This simplification is at the cost of two-dimensional Fourier transforms that are required to get the expressions in wavenumber domain.

The time averaged intensity of a sound field is given by the product of pressure and velocity (2.45). As the velocity is not uniform over the surface we perform a surface integration over the vibrating area to get the total radiated power

StartLayout 1st Row 1st Column normal upper Pi 2nd Column equals double-integral Subscript negative normal infinity Superscript normal infinity Baseline one-half upper R e left-parenthesis bold-italic p left-parenthesis bold r right-parenthesis bold-italic v Subscript z Superscript asterisk Baseline left-parenthesis bold r right-parenthesis right-parenthesis d bold r 2nd Row 1st Column Blank 2nd Column equals double-integral Subscript negative normal infinity Superscript normal infinity Baseline double-integral Subscript negative normal infinity Superscript normal infinity Baseline StartFraction j omega rho 0 Over 4 pi l EndFraction upper R e left-parenthesis e Superscript minus j k l Baseline bold-italic v Subscript z Baseline left-parenthesis bold r 0 right-parenthesis bold-italic v Subscript z Superscript asterisk Baseline left-parenthesis bold r right-parenthesis right-parenthesis d bold r 0 d bold r with l equals StartAbsoluteValue ModifyingAbove r With right-arrow minus ModifyingAbove r With right-arrow Subscript 0 Baseline EndAbsoluteValue EndLayout  (2.160)

Thus, for the determination of radiated power a double area integral is required that may become computationally expensive.

In the above expression we can also switch to the wavenumber domain. In this case the area integration is replaced by an integration over the two-dimensional wavenumber space.

StartLayout 1st Row 1st Column normal upper Pi 2nd Column equals double-integral Subscript negative normal infinity Superscript normal infinity Baseline one-half upper R e left-parenthesis bold-italic p left-parenthesis bold k right-parenthesis bold-italic v Subscript z Superscript asterisk Baseline left-parenthesis bold k right-parenthesis right-parenthesis d bold k 2nd Row 1st Column Blank 2nd Column equals double-integral Subscript negative normal infinity Superscript normal infinity Baseline one-half upper R e left-parenthesis StartFraction 1 Over 4 pi EndFraction StartFraction rho 0 omega Over StartRoot k Subscript a Superscript 2 Baseline minus k squared EndRoot EndFraction bold-italic v Subscript z Baseline left-parenthesis bold k right-parenthesis bold-italic v Subscript z Superscript asterisk Baseline left-parenthesis bold k right-parenthesis right-parenthesis d bold k 3rd Row 1st Column Blank 2nd Column equals double-integral Subscript negative normal infinity Superscript normal infinity Baseline one-half upper R e left-parenthesis StartFraction 1 Over 4 pi EndFraction StartFraction rho 0 omega Over StartRoot k Subscript a Superscript 2 Baseline minus k squared EndRoot EndFraction v Subscript z Superscript 2 Baseline left-parenthesis bold k right-parenthesis right-parenthesis d bold k EndLayout  (2.161)

The double integral is replaced by a single two-dimensional wavenumber integration. Thus, once the shape function is available the power calculation in wavenumber space is much faster than in real space (Graham, 1996).

2.7.4.1 Radiation Efficiency

The radiation efficiency is a quantity that relates the power of a plane wave to the radiated power of a vibrating surface with same surface averaged velocity. The definition of the radiation efficiency was motivated by experimental procedures because it allows the estimation of the radiated power from the measurements of the vibration velocity. The squared average velocity of a vibrating surface is

mathematical left-angle ModifyingAbove v With caret Subscript z Superscript 2 Baseline mathematical right-angle Subscript upper S Baseline equals StartFraction 1 Over upper S EndFraction integral Underscript upper S Endscripts bold-italic v Subscript z Baseline left-parenthesis bold r right-parenthesis bold-italic v Subscript z Superscript asterisk Baseline left-parenthesis bold r right-parenthesis d bold r  (2.162)

and the power radiated by a plane wave through the same area S is given by (2.47)

normal upper Pi 0 equals one-half upper S rho 0 c 0 mathematical left-angle ModifyingAbove v With caret Subscript z Superscript 2 Baseline mathematical right-angle Subscript upper S  (2.163)

The radiation efficiency is defined as the ratio between the radiated power of a velocity profile vz(r) of a surface S and the standardized power of the plane wave:

sigma Subscript normal r normal a normal d Baseline equals StartFraction normal upper Pi Over normal upper Pi 0 EndFraction equals StartStartFraction normal upper Pi OverOver StartFraction upper S Over 2 EndFraction rho 0 c 0 mathematical left-angle ModifyingAbove v With caret Subscript z Superscript 2 Baseline mathematical right-angle Subscript upper S Baseline EndEndFraction  (2.164)

The radiation efficiency is used to determine the radiated power of vibrating structures from calculated, estimated, or measured radiation efficiency of specific surfaces

normal upper Pi equals StartFraction upper S Over 2 EndFraction sigma Subscript normal r normal a normal d Baseline rho 0 c 0 mathematical left-angle ModifyingAbove v With caret Subscript z Superscript 2 Baseline mathematical right-angle Subscript upper S Baseline equals upper S sigma Subscript normal r normal a normal d Baseline rho 0 c 0 mathematical left-angle v Subscript z comma rms Superscript 2 Baseline mathematical right-angle Subscript upper S  (2.165)

2.8 Units, Measures, and levels

The dynamic range of acoustic quantities can be very high; thus, a logarithmic scale is well established for the quantification of acoustic signals. For convenience a certain time averaged quantity, for example the mean square pressure pT2=prms2, is compared to a mean square reference value pref2. The pressure level in decibels is defined as follows:

upper L Subscript p Baseline equals 10 log Subscript 10 Baseline StartFraction p Subscript normal r normal m normal s Superscript 2 Baseline Over p Subscript ref Superscript 2 Baseline EndFraction equals 20 log Subscript 10 Baseline StartFraction p Subscript normal r normal m normal s Baseline Over p Subscript ref Baseline EndFraction with p Subscript ref Baseline equals 20 mu Pa  (2.166)

The factor of 10 is introduced to spread the scale. Linear quantities such as pressure, velocity, or displacement use the mean square values. As level and decibel are used on time signals too, one should not apply the decibel scale to amplitudes. This may lead to confusion, as it is not clear if the mean square values of the amplitude is meant. This makes even more sense when the energy and power levels are defined. The energy must be averaged, as there is no constant value over the period – see Equation (2.46). Energy quantities such as energy, intensity, or power are compared with mean values and not mean square values; hence:

upper L Subscript w Baseline equals 10 log Subscript 10 Baseline StartFraction normal upper Pi Subscript mean Baseline Over normal upper Pi Subscript ref Baseline EndFraction with normal upper Pi Subscript ref Baseline equals 10 Superscript negative 12 Baseline normal upper W equals 1 pW  (2.167)

Table 2.3Field and energy properties of acoustic waves

SourceSource strengthImpedance VelocityPressure Radiated power
Mono poleQ,jωVZrad=k2ρ0c04πp=jkρ0c0Q(ω)4πrejkr
vr=Q(ω)4πrjk(1+1jkr)ejkrQrms2k2ρ0c04π
Breath. sphereQ=4πR2vrzR=jρ0c0kR1+jkRQ4πr[jkρ0c01+jkR]ejk(rR)
Q4πr2[1+jkr1+jkR]ejk(rR)Qrms2k2ρ0c04π(1+k2R2)
PistonQ=πR2vzzR=ρ0c0(1J1(2kR)kR+jH1(2kR)kR)jωρ02πrvz[2J1(kRsinϑ)kRsinϑ]
ρ0c0(1J1(2kR)kR+jH1(2kR)kR)Qrms2

In addition, the decibel scale is used for ratios of similar quantities. A typical example is the transmission loss that is the decibel scale of the transmission factor from Equation (2.118) that relates the transmitted to the radiated power. The definition of the transmission loss (TL) is:

TL equals 10 log Subscript 10 Baseline StartFraction 1 Over tau EndFraction equals minus 10 log Subscript 10 Baseline tau  (2.168)

The reciprocal definition was chosen in order to get positive values for losses. When linear quantities are compared, for example the pressure at two locations, the mean square values are related. When the squared pressure is compared to the situation with and without a specific equipment or installation, this is called insertion loss (IL)

IL equals 10 log Subscript 10 Baseline StartFraction p Subscript out Superscript 2 Baseline Over p Subscript in Superscript 2 Baseline EndFraction equals 20 log Subscript 10 Baseline StartFraction p Subscript out Baseline Over normal p Subscript in Baseline EndFraction  (2.169)

The reference quantities for power and pressure are chosen conveniently to simplify the calculations with levels. Taking the equation for the spherical source (2.82) and dividing it by the squared reference value for the pressure yields

StartFraction p Subscript normal r normal m normal s Superscript 2 Baseline Over p Subscript ref Superscript 2 Baseline EndFraction equals StartFraction rho 0 c 0 Over p Subscript ref Superscript 2 Baseline upper A Subscript ref Baseline EndFraction StartFraction upper A Subscript ref Baseline Over 4 pi r squared EndFraction mathematical left-angle normal upper Pi mathematical right-angle Subscript upper T  (2.170)

and taking the decibel of this

StartLayout 1st Row 1st Column 10 log Subscript 10 Baseline left-brace StartFraction p Subscript normal r normal m normal s Superscript 2 Baseline Over p Subscript ref Superscript 2 Baseline EndFraction right-brace 2nd Column equals 10 log Subscript 10 Baseline left-brace StartStartFraction mathematical left-angle normal upper Pi mathematical right-angle Subscript upper T Baseline OverOver StartFraction upper A Subscript ref Baseline p Subscript ref Superscript 2 Baseline Over rho 0 c 0 EndFraction EndEndFraction right-brace plus 10 log Subscript 10 Baseline left-brace StartFraction upper A Subscript ref Baseline Over 4 pi r squared EndFraction right-brace EndLayout

yields

StartLayout 1st Row 1st Column upper L Subscript p 2nd Column equals upper L Subscript w Baseline plus 10 log Subscript 10 Baseline left-brace StartFraction upper A Subscript ref Baseline Over 4 pi r squared EndFraction right-brace with normal upper Pi prime Subscript ref Baseline equals StartFraction upper A Subscript ref Baseline p Subscript ref Superscript 2 Baseline Over rho 0 c 0 EndFraction period EndLayout  (2.171)

Entering typical values for air with ρ0=1.23 kg/m 3 and c0=343 m/s using Aref=1 m 2 we get

normal upper Pi prime Subscript ref Baseline almost-equals 10 Superscript negative 12 Baseline normal upper W equals normal upper Pi Subscript ref

matching well to the reference value of acoustic power.

Bibliography

  1. W.R. Graham. BOUNDARY LAYER INDUCED NOISE IN AIRCRAFT, PART I: THE FLAT PLATE MODEL. Journal of Sound and Vibration, 192(1): 101–120, April 1996. ISSN 0022-460X.
  2. Finn Jacobsen. PROPAGATION OF SOUND WAVES IN DUCTS. Technical Note 31260, Technical University of Denmark, Lynby, Denmark, September 2011.
  3. Reinhard Lerch and H. Landes. Grundlagen der Technischen Akustik, September 2012.
  4. P.M.C. Morse and K.U. Ingard Theoretical Acoustics. International Series in Pure and Applied Physics. Princeton University Press, 1968. ISBN 978-0-691-02401-1.

Notes

  1. 1   Keeping v00 would lead to the convective wave equation that is used in the context of flow related acoustic problems, which is more a topic of aero acoustics
  2. 2   Negative k means propagation in positive x-direction, because of the jωt convention. This is the reason why pure acoustic textbooks usually take the jωt convention in order to get positive wavenumber for positive propagation.
  3. 3   This can be proven using the relationship between the δ and exponential functions
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset