9
Deterministic Applications

In technical acoustics deterministic systems are usually treated by numeric methods like the finite element method or the boundary element method. Those methods require complex and powerful solvers as far as pre and post-processors to handle large and detailed models. Even though such models are extremely useful for the simulation of vibroacoustic systems, it is hard to develop a deep understanding of the dynamic phenomena with these numeric models and to draw the right conclusions.

In this book we will treat deterministic systems as far as possible by analytical approaches or by models that consist of sub-elements that can be described by analytical formulas. This allows the reader to follow and understand the details of the theory and may help to provide a deeper understanding of typical vibroacoustic systems. However, even with such constraints, the examples in this chapter are about several deterministic subsystems that are used in real technical systems and create the basement for later SEA or hybrid FEM/SEA examples.

9.1 Acoustic One-Dimensional Elements

One-dimensional acoustic elements are used in the simulation of mufflers, ventilation systems, or hydraulics. The wavelength is assumed to be much larger than the dimension of the cross section. Such systems can become very complex, and they are also used as a designed network in audio systems, for example, the housing and resonators of loudspeakers. In the context of this book, these elements are presented as typical deterministic applications in order to explain the different effects of filters, resonators, and absorbers.

9.1.1 Transfer Matrix and Finite Element Convention

When dealing with one-dimensional systems, the literature often refers to the transfer matrix theory (Pierce, 1991; Mechel, 2002). This approach is useful when the full system is also one-dimensional, meaning that it is a linear chain of subsystems without additional branches.

In Figure 9.1 the convention for both is shown. When the pressure (or the velocity) is used as state variable the mobility matrix reads as (4.11)

Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper Y 11 2nd Column bold-italic upper Y 12 2nd Row 1st Column bold-italic upper Y 21 2nd Column bold-italic upper Y 22 EndMatrix StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix equals StartBinomialOrMatrix bold-italic v Subscript x Baseline Subscript 1 Baseline Choose bold-italic v Subscript x Baseline Subscript 2 EndBinomialOrMatrix  (9.1)

Figure 9.1Convention for stiffness matrix and TMM. Source: Alexander Peiffer.

or the impedance matrix

Start 2 By 2 Matrix 1st Row 1st Column bold-italic z 11 2nd Column bold-italic z 12 2nd Row 1st Column bold-italic z 21 2nd Column bold-italic z 22 EndMatrix StartBinomialOrMatrix bold-italic v Subscript x Baseline Subscript 1 Baseline Choose bold-italic v Subscript x Baseline Subscript 2 Baseline EndBinomialOrMatrix equals StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix  (9.2)

The velocity in equation (9.2) is a shared internal degree of freedom, and the pressure corresponds to an external pressure as discussed in Xue (2003). Thus, due to the continuity of pressure (or force), we assume for the external pressure p2L=p2R on the right hand side.

In the transfer matrix method theory the situation is different – here the pressure is the internal pressure or the state variable, and pint,2L=pint,2R.

To conclude, when we switch from the transfer matrix method presentation to FE, the right hand side internal pressure is the negative internal pressure: p2R=pint,2R. In the following we will use p2 in the mobility presentation and p2 in the transfer matrix approach. The transfer matrix representation of the same system is:

StartBinomialOrMatrix bold-italic p 1 Choose bold-italic v Subscript x Baseline Subscript 1 Baseline EndBinomialOrMatrix equals Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper T 11 2nd Column bold-italic upper T 12 2nd Row 1st Column bold-italic upper T 21 2nd Column bold-italic upper T 22 EndMatrix StartBinomialOrMatrix bold-italic p prime 2 Choose bold-italic v Subscript x Baseline Subscript 2 Baseline EndBinomialOrMatrix equals StartBinomialOrMatrix minus bold-italic p 2 Choose bold-italic v Subscript x Baseline Subscript 2 EndBinomialOrMatrix  (9.3)

Both representations can be easily exchanged. Solving the above equation for each different state variable gives

StartLayout 1st Row 1st Column StartFraction 1 Over bold-italic upper T 12 EndFraction Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper T 22 2nd Column def left-parenthesis Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix right-parenthesis 2nd Row 1st Column 1 2nd Column bold-italic upper T 11 EndMatrix StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix 2nd Column equals StartBinomialOrMatrix bold-italic v Subscript x Baseline Subscript 1 Baseline Choose bold-italic v Subscript x Baseline Subscript 2 EndBinomialOrMatrix EndLayout  (9.4)
StartLayout 1st Row StartBinomialOrMatrix bold-italic p 1 Choose bold-italic v Subscript x Baseline Subscript 1 Baseline EndBinomialOrMatrix equals StartFraction 1 Over bold-italic upper Y 21 EndFraction Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper Y 22 2nd Column 1 2nd Row 1st Column def Start 1 By 1 Matrix 1st Row bold-italic upper Y EndMatrix 2nd Column bold-italic upper Y 11 EndMatrix StartBinomialOrMatrix bold-italic p prime 2 Choose bold-italic v Subscript x Baseline Subscript 2 EndBinomialOrMatrix EndLayout  (9.5)

When the system consists of a cascade of one-dimensional systems, the transfer matrix method is very convenient

StartBinomialOrMatrix bold-italic p 1 Choose bold-italic v 1 EndBinomialOrMatrix equals Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript 12 Baseline StartBinomialOrMatrix bold-italic p 2 Choose bold-italic v 1 EndBinomialOrMatrix StartBinomialOrMatrix bold-italic p 2 Choose bold-italic v 2 EndBinomialOrMatrix equals Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript 23 Baseline StartBinomialOrMatrix bold-italic p 3 Choose bold-italic v 3 EndBinomialOrMatrix ellipsis StartBinomialOrMatrix bold-italic p Subscript upper N minus 1 Baseline Choose bold-italic v Subscript upper N minus 1 Baseline EndBinomialOrMatrix equals Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript upper N minus 1 comma upper N Baseline StartBinomialOrMatrix bold-italic p Subscript upper N Baseline Choose bold-italic v Subscript upper N EndBinomialOrMatrix  (9.6)

because the total transfer matrix is the product of all transfer matrices

Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript 1 upper N Baseline equals Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript 12 Baseline Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript 23 Baseline midline-horizontal-ellipsis Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript upper N minus 1 comma upper N  (9.7)

This makes the calculation fast and simple, because no matrix inversion is involved. However, the FE approach is more straightforward and allows for branches in the total system.

9.1.2 Acoustic One-Dimensional Networks

Acoustic networks consist of systems with specific cross sections Ac(i) at both ends. Thus, it is useful in accordance with the finite element formulation from section 4.3.2 to switch from velocity to volume flow. This is the continuous quantity at connections as shown in Figure 9.2

Figure 9.2Connection of one-dimensional acoustic systems of different cross sections. Source: Alexander Peiffer.

The appropriate matrix equation to describe this would be the radiation mobility matrix with the pressure as internal state variable and the volume flow as external excitation quantity.

Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper Y Subscript a comma 11 Baseline 2nd Column bold-italic upper Y Subscript a comma 12 Baseline 2nd Row 1st Column bold-italic upper Y Subscript a comma 21 Baseline 2nd Column bold-italic upper Y Subscript a comma 22 Baseline EndMatrix StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix equals StartBinomialOrMatrix bold-italic upper Q 1 Choose bold-italic upper Q 2 EndBinomialOrMatrix  (9.8)

An example for an acoustic network is shown in Figure 9.2. Here the net flow into the nodes is zero when no external volume source is applied. In the final matrix equation, the state variables Qi are calculated and must be derived from the impedance of the connected and cut-free subsystems. So, the flow Qi into each node i determines the nodal pressure as state solution. We use the following convention: The flow Qin denotes the volume flow into the node i from the system n.

The equation of motion for the acoustic system as shown in Figure 9.3 is

Start 5 By 5 Matrix 1st Row 1st Column bold-italic upper Y Subscript a comma 11 Baseline 2nd Column bold-italic upper Y Subscript a comma 12 Baseline 3rd Column midline-horizontal-ellipsis 4th Column Blank 5th Column bold-italic upper Y Subscript a comma 15 Baseline 2nd Row 1st Column bold-italic upper Y Subscript a comma 21 Baseline 2nd Column bold-italic upper Y Subscript a comma 22 Baseline 3rd Column midline-horizontal-ellipsis 4th Column Blank 5th Column bold-italic upper Y Subscript a comma 25 Baseline 3rd Row 1st Column vertical-ellipsis 2nd Column Blank 3rd Column down-right-diagonal-ellipsis 4th Column Blank 5th Column Blank 4th Row 1st Column vertical-ellipsis 2nd Column Blank 3rd Column Blank 4th Column down-right-diagonal-ellipsis 5th Column Blank 5th Row 1st Column bold-italic upper Y Subscript a comma 51 Baseline 2nd Column bold-italic upper Y Subscript a comma 42 Baseline 3rd Column Blank 4th Column midline-horizontal-ellipsis 5th Column bold-italic upper Y Subscript a comma 55 Baseline EndMatrix Start 5 By 1 Matrix 1st Row bold-italic p 1 2nd Row bold-italic p 2 3rd Row bold-italic p 3 4th Row bold-italic p 4 5th Row bold-italic p 5 EndMatrix equals Start 5 By 1 Matrix 1st Row bold-italic upper Q Subscript 1 comma ext Baseline 2nd Row 0 3rd Row 0 4th Row 0 5th Row 0 EndMatrix period  (9.9)

Figure 9.3Acoustic network with nodal volume flow, volume sources, and radiation impedance at open ends. The numbers in circles denote the system numbers. Source: Alexander Peiffer.

Equation (9.9) is derived by using the element mobility from (9.8) and adding each element mobility to the total system matrix similar to the procedure described in section 4.3.1 but with a different source term.

This finite element formulation is efficient when used as a numerical solution but not when analytical expressions are required. The network equation must be solved or inverted to get the system response. Inverting the analytical formulas is not easily done or possible in many cases, and the transfer matrices are more useful.

According to the discussions regarding the transfer matrices in section 9.1.1, we can change easily between the different formulations. The transfer matrix [Ta] reads as:

StartBinomialOrMatrix bold-italic p 1 Choose bold-italic upper Q 1 EndBinomialOrMatrix equals Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper T Subscript a comma 11 Baseline 2nd Column bold-italic upper T Subscript a comma 12 Baseline 2nd Row 1st Column bold-italic upper T Subscript a comma 21 Baseline 2nd Column bold-italic upper T Subscript a comma 22 Baseline EndMatrix StartBinomialOrMatrix bold-italic p prime 2 Choose bold-italic upper Q 2 EndBinomialOrMatrix equals StartBinomialOrMatrix minus bold-italic p 2 Choose bold-italic upper Q 2 EndBinomialOrMatrix  (9.10)

and the conversion between each representation is

StartLayout 1st Row 1st Column StartFraction 1 Over upper T Subscript a comma 12 Baseline EndFraction 2nd Column Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper T Subscript a comma 22 Baseline 2nd Column def left-parenthesis Start 1 By 1 Matrix 1st Row bold-italic upper T Subscript a Baseline EndMatrix right-parenthesis 2nd Row 1st Column 1 2nd Column bold-italic upper T Subscript a comma 11 Baseline EndMatrix StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix equals StartBinomialOrMatrix bold-italic upper Q 1 Choose bold-italic upper Q 2 EndBinomialOrMatrix EndLayout  (9.11)
StartLayout 1st Row 1st Column StartBinomialOrMatrix bold-italic p 1 Choose bold-italic upper Q 1 EndBinomialOrMatrix equals StartFraction 1 Over upper Y Subscript a comma 21 Baseline EndFraction 2nd Column Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper Y Subscript a comma 22 2nd Column 1 2nd Row 1st Column def Start 1 By 1 Matrix 1st Row bold-italic upper Y Subscript a Baseline EndMatrix 2nd Column bold-italic upper Y Subscript a comma 11 EndMatrix StartBinomialOrMatrix bold-italic p prime 2 Choose bold-italic upper Q 2 EndBinomialOrMatrix EndLayout  (9.12)

Thus, for one-dimensional acoustic networks with changing cross section, equation (9.10) may be the best choice.

9.1.2.1 Properties of the System Matrices

From the reciprocity principle some useful properties can be derived. Reciprocity states

StartFraction bold-italic upper Q 1 Over bold-italic p 2 EndFraction equals StartFraction bold-italic upper Q 2 Over bold-italic p 1 EndFraction  (9.13)

Entering this into the two port equation gives for this mobility matrix

StartLayout 1st Row 1st Column bold-italic upper Y Subscript a comma 11 2nd Column equals bold-italic upper Y Subscript a comma 22 Baseline 3rd Column bold-italic upper Y Subscript a comma 12 4th Column equals bold-italic upper Y Subscript a comma 21 EndLayout  (9.14)

Hence, they are symmetric. The transfer matrix is obviously not symmetric which can be seen from equations (9.5) and (9.12). From the same equations it can be derived that the determinant of the transfer matrix equals 1

def Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix equals 1  (9.15)

9.1.3 The Acoustic Pipe

Figure 9.4Properties of an acoustic pipe. Source: Alexander Peiffer.

We take the acoustic pipe from section 4.1 and switch to the volume flow using Q=Acv; we get the radiation mobility matrix formulation with pressure as state variable and the volume flow as external source from equation (4.11):

StartFraction upper A Subscript c Baseline Over bold-italic z EndFraction Start 2 By 2 Matrix 1st Row 1st Column StartFraction 1 Over j tangent left-parenthesis bold-italic k upper L right-parenthesis EndFraction 2nd Column StartFraction 1 Over j sine left-parenthesis bold-italic k upper L right-parenthesis EndFraction 2nd Row 1st Column StartFraction 1 Over j sine left-parenthesis bold-italic k upper L right-parenthesis EndFraction 2nd Column StartFraction 1 Over j tangent left-parenthesis bold-italic k upper L right-parenthesis EndFraction EndMatrix StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix equals StartBinomialOrMatrix bold-italic upper Q 1 Choose bold-italic upper Q 2 EndBinomialOrMatrix  (9.16)

The transfer matrix representation reads as:

StartBinomialOrMatrix bold-italic p 1 Choose bold-italic upper Q 1 EndBinomialOrMatrix equals Start 2 By 2 Matrix 1st Row 1st Column cosine left-parenthesis bold-italic k upper L right-parenthesis 2nd Column j StartFraction rho 0 c 0 Over upper A Subscript c Baseline EndFraction sine left-parenthesis bold-italic k upper L right-parenthesis 2nd Row 1st Column j StartFraction upper A Subscript c Baseline Over rho 0 c 0 EndFraction sine left-parenthesis bold-italic k upper L right-parenthesis 2nd Column cosine left-parenthesis bold-italic k upper L right-parenthesis EndMatrix StartBinomialOrMatrix bold-italic p prime 2 Choose bold-italic upper Q 2 EndBinomialOrMatrix  (9.17)

9.1.4 Volumes and Closed Pipes

Figure 9.5 Closed volume and pipe. Source: Alexander Peiffer.

The closed volume was derived in section 4.3.1 leading to equation (4.98). We divide this equation by jω in order to get the mobility version with Q as source term

j omega StartFraction upper V Over upper K EndFraction bold-italic p equals bold-italic upper Q  (9.18)

The bulk modulus can be replaced using (2.18), giving

j omega StartFraction upper V Over rho 0 c 0 squared EndFraction bold-italic p equals bold-italic upper Q  (9.19)

The volume formulation assumes a volume extension to be much smaller than the wavelength. For thin cylinder shaped volumes, the one-dimensional pipe formulation with rigid end (4.16) leads to

j StartFraction upper A Subscript c Baseline tangent left-parenthesis bold-italic k upper L right-parenthesis Over z 0 EndFraction bold-italic p equals bold-italic upper Q  (9.20)

This equation is more appropriate for volumes that have only cross section dimensions that are small compared to the wavelength but can have large dimensions in the direction of sound propagation. However, when kL=2πωLc01, the tangent can be approximated by tanxx, and equation (9.20) leads to

j StartFraction upper A Subscript c Baseline k upper L Over z 0 EndFraction bold-italic p equals j omega StartFraction upper A Subscript c Baseline upper L Over rho 0 c 0 squared EndFraction bold-italic p equals bold-italic upper Q  (9.21)

This is exactly corresponding to expression (9.19). We conclude with the mobility of the volume and tube, namely

StartLayout 1st Row 1st Column volume colon bold-italic upper Y Subscript a comma 1 upper D Baseline 2nd Column equals bold-italic p equals j omega StartFraction upper A Subscript c Baseline upper L Over rho 0 c 0 squared EndFraction 3rd Column pipe colon bold-italic upper Y Subscript a comma 1 upper D Baseline 4th Column equals j StartFraction upper A Subscript c Baseline tangent left-parenthesis bold-italic k upper L right-parenthesis Over z 0 EndFraction EndLayout  (9.22)

and the according impedances from the reciprocal.

9.1.5 Limp Layer

This generic model describes a lumped element in the pipe flow, representing mass, stiffness, or damping effects. The condition for the validity of lumped elements is that the wavenumber must be much larger than the dimension of the element. Thus, we assume v1=v2.

As shown in Figure 9.6, the lumped element presentation has in common that the volume flow or velocity is equal on both sides. Thus, the dynamic behavior can be described by transfer impedances

StartLayout 1st Row 1st Column bold-italic z Subscript 1 upper D 2nd Column equals StartFraction bold-italic p prime 2 minus bold-italic p prime 1 Over bold-italic v EndFraction equals StartFraction normal upper Delta bold-italic p Over bold-italic v EndFraction 3rd Column bold-italic upper Z Subscript a comma 1 upper D 4th Column equals StartFraction bold-italic p prime 2 minus bold-italic p prime 1 Over upper A Subscript c Baseline bold-italic v EndFraction equals StartFraction normal upper Delta bold-italic p Over bold-italic upper Q EndFraction EndLayout  (9.23)

Figure 9.6 Mass and stiffness element in a pipe of cross setion Ac. Source: Alexander Peiffer.

The transfer impedance can be written in the following way:

StartLayout 1st Row 1st Column bold-italic z Subscript 1 upper D 2nd Column equals upper R Subscript 1 upper D Baseline plus j upper X Subscript 1 upper D Baseline 3rd Column bold-italic upper Z Subscript a comma 1 upper D 4th Column equals upper R Subscript a comma 1 upper D Baseline plus j upper X Subscript a comma 1 upper D Baseline 5th Column with bold-italic upper Z Subscript a comma 1 upper D Baseline 6th Column equals StartFraction bold-italic z Subscript 1 upper D Baseline Over upper A Subscript c Baseline EndFraction EndLayout  (9.24)

The real and imaginary parts represent the reactive and dissipative parts, respectively. The transfer matrix of such an element is then determined by reshuffling the above equations to an appropriate set

StartLayout 1st Row 1st Column bold-italic p prime 1 2nd Column equals bold-italic p prime 2 plus bold-italic upper Z Subscript 1 upper D Baseline bold-italic v 2 EndLayout  (9.25)
StartLayout 1st Row 1st Column bold-italic v 1 2nd Column equals bold-italic v 2 EndLayout  (9.26)

and the transfer matrix of the generic layer reads as:

bold-italic upper T Subscript 1 upper D Baseline equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column bold-italic z Subscript 1 upper D Baseline 2nd Row 1st Column 0 2nd Column 1 EndMatrix bold-italic upper T Subscript a comma 1 upper D Baseline equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column bold-italic upper Z Subscript a comma 1 upper D Baseline 2nd Row 1st Column 0 2nd Column 1 EndMatrix  (9.27)

According to the transformation in (9.11) the radiation mobility matrix representation is

StartFraction 1 Over bold-italic upper Z Subscript a comma 1 upper D Baseline EndFraction Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column 1 2nd Row 1st Column 1 2nd Column 1 EndMatrix StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix equals StartBinomialOrMatrix bold-italic upper Q 1 Choose bold-italic upper Q 2 EndBinomialOrMatrix  (9.28)

In the literature the transfer matrix is often used to describe the acoustics of specific layers.

9.1.5.1 Mass

Mass layers can be limp membranes that are closing the pipe or plates that are in the pipe where friction can be neglected. For example, a thin fluid layer of small thickness can also be approximated by a mass layer. Following Newton’s law F=mx¨, the equation of motion is

upper A Subscript c Baseline left-parenthesis bold-italic p prime 1 minus bold-italic p prime 2 right-parenthesis equals j omega m bold-italic v long left right double arrow bold-italic z Subscript 1 upper D Baseline equals StartFraction j omega m Over upper A Subscript c Baseline EndFraction equals j omega m double-prime  (9.29)

Similar to the discussion concerning volume and pipe end the mobility expressions can also be derived as an approximation of equation (9.16) for small values of |kL|.

9.1.5.2 Stiffness

A stiffness in the pipe can be thought of as an infinitely stiff plate supported by a spring of stiffness ks or a specific stiffness ks=ks/Ac. The equation of motion for a spring F=kx leads to

upper A Subscript c Baseline left-parenthesis bold-italic p prime 1 minus bold-italic p prime 2 right-parenthesis equals minus k Subscript s Baseline StartFraction bold-italic v Over j omega EndFraction right double arrow bold-italic z Subscript 1 upper D Baseline equals minus StartFraction k Subscript s Baseline Over j omega upper A Subscript c Baseline EndFraction equals minus StartFraction k double-prime Subscript s Baseline Over j omega EndFraction  (9.30)

Figure 9.7Cylindrical membrane exposed to pressure. Source: Alexander Peiffer.

9.1.5.3 Viscous Damping

The viscous damping is determined by F=cvv, and hence

upper A Subscript c Baseline left-parenthesis bold-italic p 1 minus p prime 2 right-parenthesis equals minus c Subscript v Baseline bold-italic v right double arrow bold-italic z Subscript 1 upper D Baseline equals minus StartFraction c Subscript v Baseline Over upper A Subscript c Baseline EndFraction equals minus c Subscript v Superscript double-prime Baseline equals upper R Subscript 1 upper D  (9.31)

We see that the damping results in the real resistance R, whereas the imaginary part considers mass and stiffness effects X. The total transfer impedance is hence

bold-italic z Subscript 1 upper D Baseline equals c double-prime Subscript v Baseline plus j left-parenthesis omega m double-prime plus StartFraction k double-prime Subscript s Baseline Over omega EndFraction right-parenthesis  (9.32)

9.1.6 Membranes

A practical implementation of a system with mainly mass and stiffness is a membrane with tension T0 mounted in a circular cross section Ac. From the membrane equation of motion (3.139)

upper T 0 normal upper Delta w equals upper P 1 minus upper P 2  (9.33)

and using the cylindrical Laplace operator Δ=1rr+2r2 we get

StartFraction 1 Over r EndFraction StartFraction partial-differential w Over partial-differential r EndFraction plus StartFraction partial-differential squared w Over partial-differential r squared EndFraction equals upper P 1 minus upper P 2 equals normal upper Delta upper P  (9.34)

When we assume a circular membrane with boundary condition w(R)=0, the solution is

w left-parenthesis r right-parenthesis equals w 0 left-parenthesis 1 minus StartFraction r squared Over upper R squared EndFraction right-parenthesis with w 0 equals StartFraction normal upper Delta upper P upper R squared Over 4 upper T 0 EndFraction  (9.35)

When we derive the limp parameter, we have to use the surface averaged quantities. The stiffness ks can be defined based on the volume change due to the static pressure:

k Subscript s Baseline equals minus StartFraction upper F Over w Subscript eff Baseline EndFraction equals minus StartFraction upper A Subscript c Baseline normal upper Delta upper P Over w Subscript eff Baseline EndFraction equals minus StartFraction upper A Subscript c Superscript 2 Baseline normal upper Delta upper P Over upper V EndFraction with w Subscript eff Baseline equals upper V slash upper A Subscript c Baseline  (9.36)

The area specific stiffness is defined by ks=AcΔP/V and the volume is given by

upper V equals 2 pi integral Subscript 0 Superscript upper R Baseline w left-parenthesis r right-parenthesis r d r equals 2 pi integral Subscript 0 Superscript upper R Baseline w 0 left-parenthesis 1 minus StartFraction r squared Over upper R squared EndFraction right-parenthesis r d r equals w 0 StartFraction pi upper R squared Over 2 EndFraction equals StartFraction pi upper R squared Over 8 upper T 0 EndFraction normal upper Delta upper P  (9.37)

Thus, the stiffness of the membrane is

k double-prime Subscript s Baseline equals 8 upper T 0  (9.38)

The membrane is displaced non-uniformly. For the efficient mass estimation we use the displacement weff=V/Ac=w0/2 and calculate the kinetic energy based on this

upper E Subscript kin Baseline equals StartFraction omega squared Over 2 EndFraction m Subscript eff Baseline w Subscript eff Superscript 2 Baseline equals StartFraction omega Over 2 EndFraction m Subscript eff Baseline StartFraction w 0 Over 4 EndFraction  (9.39)

The velocity maximum occurs at zero displacement position

v Subscript z Baseline left-parenthesis r right-parenthesis equals upper R e left-parenthesis j omega bold-italic w left-parenthesis r right-parenthesis e Superscript j omega t Baseline right-parenthesis equals omega w left-parenthesis r right-parenthesis

This energy must be equal to the kinetic energy of the membrane movement integrated over the given velocity shape

StartLayout 1st Row 1st Column upper E Subscript kin 2nd Column equals one-half rho 0 h Baseline 2 pi integral Subscript 0 Superscript upper R Baseline ModifyingAbove v With caret Subscript z Baseline left-parenthesis r right-parenthesis squared r d r equals one-half rho 0 h Baseline 2 pi omega squared integral Subscript 0 Superscript upper R Baseline w 0 squared left-parenthesis 1 minus StartFraction r squared Over upper R squared EndFraction right-parenthesis r d r 2nd Row 1st Column Blank 2nd Column equals rho 0 h omega squared w 0 squared StartFraction pi upper R squared Over 6 EndFraction EndLayout  (9.40)

Setting both kinetic energies equal leads to the efficient mass

m Subscript eff Baseline equals StartFraction 4 rho 0 h pi upper R squared Over 3 EndFraction  (9.41)

or the specific mass

m double-prime Subscript eff Baseline equals StartFraction 4 rho 0 h Over 3 EndFraction  (9.42)

which corresponds to 43 of a mass layer. Thus, the final reactance of the membrane is:

upper X Subscript 1 upper D comma normal m normal e normal m normal b normal r normal a normal n normal e Baseline equals four-thirds rho 0 h omega plus StartFraction 8 upper T 0 Over omega EndFraction  (9.43)

9.1.7 Perforated Sheets

Perforated sheets are a major device for acoustic treatment in engineering acoustics. Consider for example the ceiling of typical offices showing perforated surfaces or surfaces with a regular grid of holes. Such systems provide control over the parameters mass, stiffness, and damping. As a first attempt one might consider the volume in the hole as a mass with m=ρ0LShole plus a specific end correction resulting from a fluid volume partition that is moving on both ends of the channel. But, the channels are supposed to be so small that the wave motion is affected by friction at the walls, and we have a certain flow profile in the hole. A detailed treatment of this theory would go too far, but an established model is derived by Maa (1998) and extended with further details by Fuchs and Zha (1995).

In order to get more insight, only the basic concepts of Maa’s model are given here. The pore is assumed to be cylindrical and so small that friction is affecting the flow profile in the fluid. Maa derived that the velocity profile is given by

v left-parenthesis r right-parenthesis equals minus StartFraction 1 Over j omega rho 0 EndFraction StartFraction partial-differential p Over partial-differential z EndFraction left-parenthesis 1 minus StartFraction upper J 0 left-parenthesis k Subscript s h Baseline r StartRoot negative j EndRoot right-parenthesis Over upper J 0 left-parenthesis k Subscript s h Baseline upper R StartRoot negative j EndRoot right-parenthesis EndFraction right-parenthesis  (9.44)

with

k Subscript s h Baseline equals StartFraction omega rho 0 Over eta Subscript s h Baseline EndFraction  (9.45)

and ηsh as shear or dynamic viscosity. In order to get the average or efficient velocity we have to integrate over the cross section

v Subscript eff comma hole Baseline equals StartFraction 2 pi integral Subscript 0 Superscript upper R Baseline v left-parenthesis r right-parenthesis r d r Over pi upper R squared EndFraction equals minus StartFraction 1 Over j omega rho 0 EndFraction StartFraction partial-differential p Over partial-differential z EndFraction left-parenthesis 1 minus StartFraction 2 Over k Subscript s h Baseline upper R StartRoot negative j EndRoot EndFraction StartFraction upper J 1 left-parenthesis k Subscript s h Baseline upper R StartRoot negative j EndRoot right-parenthesis Over upper J 0 left-parenthesis k Subscript s h Baseline upper R StartRoot negative j EndRoot right-parenthesis EndFraction right-parenthesis  (9.46)

When we assume small hole depth h compared to the wave length we can assume pzδp/h and thus:

bold-italic z Subscript 1 upper D comma normal h normal o normal l normal e Baseline equals j omega rho 0 h left-parenthesis 1 minus StartFraction 2 Over k Subscript s h Baseline upper R StartRoot negative j EndRoot EndFraction StartFraction upper J 1 left-parenthesis k Subscript s h Baseline upper R StartRoot negative j EndRoot right-parenthesis Over upper J 0 left-parenthesis k Subscript s h Baseline upper R StartRoot negative j EndRoot right-parenthesis EndFraction right-parenthesis Superscript negative 1  (9.47)

So, we found the right expression for a single hole. The holes are covering only part of the surface. Thus, when averaging the velocity over the surface we have to consider this. The ratio of hole surface to total surface is the surface porosity σ=Sholes/S, and the efficient velocity related to the total surface S is veff=σveff, hole

and finally we get

bold-italic z Subscript 1 upper D Baseline equals StartFraction j omega rho 0 h Over sigma Superscript prime Baseline EndFraction left-parenthesis 1 minus StartFraction 2 Over k Subscript s h Baseline StartRoot negative j EndRoot EndFraction StartFraction upper J 1 left-parenthesis k Subscript s h Baseline StartRoot negative j EndRoot right-parenthesis Over upper J 0 left-parenthesis k Subscript s h Baseline StartRoot negative j EndRoot right-parenthesis EndFraction right-parenthesis Superscript negative 1 Baseline period  (9.48)

The surface porosity can be derived from the distance of the hole in a square grid as shown in Figure 9.8.

sigma prime Subscript square Baseline equals StartFraction pi upper R squared Over d squared EndFraction  (9.49)

Figure 9.8 Geometry of a perforated plate with square grid. Source: Alexander Peiffer.

In principle we have found the transfer impedance that is required to describe the dynamics of a perforated layer. But, there are some modifications required:

  1. The flow outside in the nearfield of the pore must be considered.
  2. The Bessel functions are quite unwieldy and should be simplified.

The first item is addressed by the so called end corrections. They consider the mass of the fluid above and below the pores and the additional friction at the pore edges. The second item is address by an approximation of the Bessel functions that leads to a maximal error of 5% (Fuchs and Zha, 1995).

Finally, the approximation derived by Maa and extended by the nearfield corrections in FreeFieldTechnologies (2015) is:1

StartLayout 1st Row 1st Column bold-italic upper Z Subscript 1 upper D 2nd Column equals upper R Subscript 1 upper D Baseline plus j upper X Subscript 1 upper D Baseline 2nd Row 1st Column upper R Subscript 1 upper D 2nd Column equals StartFraction 8 eta h Over sigma prime upper R squared EndFraction left-parenthesis StartRoot 1 plus StartFraction left-parenthesis k Subscript s h Baseline upper R right-parenthesis squared Over 8 EndFraction EndRoot plus StartFraction alpha StartRoot 2 EndRoot k Subscript s h Baseline upper R squared Over 8 h EndFraction right-parenthesis 3rd Row 1st Column upper X Subscript 1 upper D 2nd Column equals StartFraction omega rho 0 h Over sigma prime EndFraction left-parenthesis 1 plus StartFraction 1 Over StartRoot 9 plus StartFraction left-parenthesis k Subscript s h Baseline upper R right-parenthesis squared Over 2 EndFraction EndRoot EndFraction plus StartFraction 16 upper R Over 3 pi h EndFraction left-parenthesis 1 minus f Subscript i n t Baseline right-parenthesis right-parenthesis EndLayout  (9.50)

with α being a constant that is considered to be α=4 for sharp edged holes and α=2 for round edges; fint depends on the porosity σ and the grid pattern and is given by

f Subscript int Baseline equals left-parenthesis 1 plus epsilon right-parenthesis StartRoot sigma prime EndRoot minus epsilon StartRoot sigma Superscript prime 3 Baseline EndRoot  (9.51)

For a squared grid ϵ=0.47.

9.1.7.1 Example for a Micro Perforated Grid

The task of perforate is to provide specific acoustic mass and flow resistivity for absorber applications. In Figure 9.9 the normalized transfer impedance of a perforate with thickness h=2 mm, hole radius R=1 mm and squared grid distance d=2 mm is shown. The porosity is σ=0.0072.

Figure 9.9 Normalized transfer impedance of perforate h=2 mm, R=1 mm, d=2 mm and σ=0.0072. Source: Alexander Peiffer.

The normalized resistance is near 1 thus equal to the impedance of air. This means the perforate is well adapted to plane waves in air. The reactance shows an increase in mass by the nonlinear slope or the reactance curve resulting from viscosity effects in the pores.

9.1.8 Branch Lumped Elements

Additional to the transfer impedance, we define a branch element with radiation impedance Za,branch. The situation is contrary to the transfer impedance. As shown in Figure 9.10, the pressure is constant at both ports, but the volume flow is not.

StartLayout 1st Row 1st Column bold-italic p prime 1 2nd Column equals bold-italic p prime 2 EndLayout  (9.52)
StartLayout 1st Row 1st Column bold-italic upper Q 1 minus bold-italic upper Q 2 2nd Column equals StartFraction bold-italic p 1 Over bold-italic upper Z Subscript a comma normal b normal r normal a normal n normal c normal h Baseline EndFraction EndLayout  (9.53)

Figure 9.10 Branch impedance configuration of the pipe network. Source: Alexander Peiffer.

This leads to the transfer matrix presentation:

bold-italic upper T Subscript a comma 1 upper D Baseline equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column 0 2nd Row 1st Column 1 slash bold-italic upper Z Subscript a comma normal b normal r normal a normal n normal c normal h Baseline 2nd Column 1 EndMatrix  (9.54)

Any expression for this branch equation can be used – for example, the piston radiation impedance or the results from section 9.1.5.

9.1.9 Boundary Conditions

The boundary conditions in the network descriptions are required to define end conditions by the ratio of pressure and velocity. For the volume flow as state variable, an open end corresponds to zero pressure. When pipe systems are given, they may be excited by sources with an inner impedance, and one end might radiate into the free field. Thus, for realistic systems an impedance end condition must be defined. This could be the characteristic impedance of one-dimensional wave propagation or the radiation impedance of a piston (9.95).

The single node expression for the impedance in the finite element expression is

StartLayout 1st Row 1st Column Start 1 By 1 Matrix 1st Row bold-italic upper Y Subscript a comma i EndMatrix Start 1 By 1 Matrix 1st Row bold-italic upper Q Subscript i EndMatrix 2nd Column equals Start 1 By 1 Matrix 1st Row bold-italic p Subscript i Baseline EndMatrix 3rd Column Start 1 By 1 Matrix 1st Row bold-italic upper Y Subscript i EndMatrix Start 1 By 1 Matrix 1st Row bold-italic v Subscript i EndMatrix 4th Column equals Start 1 By 1 Matrix 1st Row bold-italic p Subscript i EndMatrix EndLayout  (9.55)

When we assume the free field boundary condition, the mobility values are

StartLayout 1st Row 1st Column bold-italic upper Y Subscript a comma i 2nd Column equals StartFraction upper A Subscript c Superscript left-parenthesis i right-parenthesis Baseline Over z 0 EndFraction 3rd Column bold-italic upper Y Subscript i 4th Column equals StartFraction 1 Over z 0 EndFraction EndLayout  (9.56)

or in case of the pipe ending in a semi infinite space, the radiating piston would be more appropriate; in this case z0 must be replaced by the impedance from equation (9.95). For unbaffled configurations the radiating sphere would be more realistic, using equation (2.84) for Za.

In case of the transfer impedance description, we set one value of the state to 1, say the velocity. In the cascade transfer matrices, the inputs of the first state vector are

StartLayout 1st Row 1st Column StartBinomialOrMatrix bold-italic p Subscript i Choose bold-italic v Subscript i EndBinomialOrMatrix 2nd Column equals StartBinomialOrMatrix bold-italic z Choose 1 EndBinomialOrMatrix 3rd Column StartBinomialOrMatrix bold-italic p Subscript i Choose bold-italic upper Q Subscript i EndBinomialOrMatrix 4th Column equals StartBinomialOrMatrix bold-italic upper Z Subscript a comma 1 Baseline Choose 1 EndBinomialOrMatrix EndLayout  (9.57)
StartLayout 1st Row 1st Column StartBinomialOrMatrix bold-italic p Subscript i Choose bold-italic v Subscript i EndBinomialOrMatrix 2nd Column equals StartBinomialOrMatrix 1 Choose bold-italic upper Y Subscript i Baseline EndBinomialOrMatrix 3rd Column StartBinomialOrMatrix bold-italic p Subscript i Choose bold-italic upper Q Subscript i EndBinomialOrMatrix 4th Column equals StartBinomialOrMatrix 1 Choose bold-italic upper Y Subscript a comma 1 EndBinomialOrMatrix EndLayout  (9.58)

9.1.10 Performance Indicators

Evaluating such networks aims at a certain effect on the noise propagation. So, we need parameters that describe the effect or performance of such networks. In general the above described systems can be set up and solved for any source and receiver configuration. There are two kinds of indicators: one is a comparative criteria given by the pressure or transmitted power with and without the device, and the other parameter is a local absorption criteria.

9.1.10.1 Transfer and Insertion Coefficients

The performance indicator results from the comparison of a reference system to the newly designed system. The ratio of squared result variables is called the insertion loss as defined in equation (2.169)

StartLayout 1st Row with Label Blank EndLabel upper I upper L 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 p Subscript in Baseline EndFraction EndLayout  (2.169)

A further quantity for system performance is the transmission coefficient.

tau equals StartFraction normal upper Pi Subscript in Baseline Over normal upper Pi Subscript out Baseline EndFraction

In Chapter 8 the transmission coefficient definition results from diffuse field reciprocity and the coupling of semi infinite systems. Equations (8.8a) and (7.27) can be used to calculate the transmission factor for a one-dimensional system by adjusting the mobility of impedance degrees of freedom to the stiffness coordinates. In the literature there are several derivations using the transfer matrix approach, for example Allard and Atalla (2009), Tageman (2013), or Jacobsen (2011). Here, the mobility matrix approach will be used.

As can be seen in Figure 9.11, an arbitrary system is extended by the radiation mobilities Ys and Yr for sender and receiver, respectively. The input power is given by

normal upper Pi Subscript in Baseline equals one-half ModifyingAbove p With caret Subscript 1 Superscript 2 Baseline upper R e bold-italic upper Y Subscript s  (9.59)

Figure 9.11Reference and test system configuration for the definition of the transmission. Source: Alexander Peiffer.

because the pressure and the (internal) volume flow are given by the entry boundary condition. The reference system is described by two connected equal mobilities Ys, so an external source Q1 leads to the pressure

bold-italic p 1 equals StartFraction bold-italic upper Q 1 Over 2 bold-italic upper Y Subscript s Baseline EndFraction  (9.60)

and finally:

normal upper Pi Subscript in Baseline equals one-half upper R e left-brace bold-italic upper Y Subscript s Baseline right-brace StartFraction upper Q 1 squared Over 4 bold-italic upper Y Subscript s Superscript 2 Baseline EndFraction  (9.61)

The power radiated to the receiver mobility Yr is:

normal upper Pi Subscript out Baseline equals one-half p 2 squared upper R e bold-italic upper Y Subscript r  (9.62)

In a complex network with many nodes, the network matrix equation (9.9) must be solved to calculate the output pressure of the network and the radiation mobilities. In order to cope with an input and output description, the network matrix must be condensed to the external degrees of freedom. Thus, the assumption of a 2x2 matrix is not a constraint of generality, and the result can be compared to the transfer matrix theory. A general one-dimensional system as shown in Figure 9.11 with radiation impedance endings is described by:

Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper Y Subscript a comma 11 Baseline plus bold-italic upper Y Subscript a comma s Baseline 2nd Column bold-italic upper Y Subscript a comma 12 Baseline 2nd Row 1st Column bold-italic upper Y Subscript a comma 21 Baseline 2nd Column bold-italic upper Y Subscript a comma 22 Baseline plus plus bold-italic upper Y Subscript a comma r Baseline EndMatrix StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix equals Start 1 By 1 Matrix 1st Row bold-italic upper Y Subscript a comma t o t Baseline EndMatrix StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix equals StartBinomialOrMatrix bold-italic upper Q 1 Choose bold-italic upper Q 2 EndBinomialOrMatrix  (9.63)

Matrix inversion gets p2 from Q1, and we consider that there is no source at port 2 hence Q2=0:

StartBinomialOrMatrix bold-italic p 1 Choose bold-italic p 2 EndBinomialOrMatrix equals StartFraction 1 Over def Start 1 By 1 Matrix 1st Row bold-italic upper Y Subscript a comma normal t normal o normal t Baseline EndMatrix EndFraction Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper Y Subscript a comma 22 Baseline plus bold-italic upper Y Subscript a comma r Baseline 2nd Column minus bold-italic upper Y Subscript a comma 21 Baseline 2nd Row 1st Column minus bold-italic upper Y Subscript a comma 12 Baseline 2nd Column bold-italic upper Y Subscript a comma 11 Baseline plus plus bold-italic upper Y Subscript a comma s Baseline EndMatrix StartBinomialOrMatrix bold-italic upper Q 1 Choose 0 EndBinomialOrMatrix  (9.64)

So, the pressure is given by

bold-italic p 2 equals StartFraction minus bold-italic upper Y Subscript a comma 12 Baseline Over d e t left-parenthesis Start 1 By 1 Matrix 1st Row bold-italic upper Y EndMatrix right-parenthesis EndFraction bold-italic upper Q 1  (9.65)

Entering this into (9.62) gives

upper Pi Subscript o u t Baseline equals one-half p 2 squared upper R e upper Y Subscript r Baseline StartFraction upper Y squared a comma 12 Over d e l left-parenthesis left-bracket upper Y right-bracket right-parenthesis squared EndFraction upper Q 1 squared  (9.66)

This reads for the transmission factor

tau equals StartFraction normal upper Pi Subscript out Baseline Over normal upper Pi Subscript in Baseline EndFraction equals 4 StartFraction upper R e bold-italic upper Y Subscript r Baseline Over upper R e bold-italic upper Y Subscript s Baseline EndFraction StartFraction bold-italic upper Y Subscript a comma 12 Superscript 2 Baseline bold-italic upper Y Subscript a comma s Superscript 2 Baseline Over StartAbsoluteValue def Start 1 By 1 Matrix 1st Row bold-italic upper Y Subscript a comma normal t normal o normal t Baseline EndMatrix EndAbsoluteValue squared EndFraction  (9.67)

When we have condensed the network, this expression defines the transmission through the network. We rearrange the details of the total matrix determinant

StartLayout 1st Row 1st Column tau 2nd Column equals StartFraction upper R e bold-italic upper Y Subscript r Baseline Over upper R e bold-italic upper Y Subscript s Baseline EndFraction StartFraction 4 bold-italic upper Y Subscript a comma 12 Superscript 2 Baseline bold-italic upper Y Subscript a comma s Superscript 2 Baseline Over StartAbsoluteValue left-bracket left-parenthesis bold-italic upper Y Subscript a comma 11 Baseline plus bold-italic upper Y Subscript a comma s Baseline right-parenthesis left-parenthesis bold-italic upper Y Subscript a comma 22 Baseline plus bold-italic upper Y Subscript a comma r Baseline right-parenthesis minus bold-italic upper Y Subscript a comma 12 Baseline plus bold-italic upper Y Subscript a comma 21 Baseline right-bracket bold-italic upper Y Subscript a comma s Baseline EndAbsoluteValue squared EndFraction 2nd Row 1st Column Blank 2nd Column equals StartFraction upper R e bold-italic upper Y Subscript r Baseline Over upper R e bold-italic upper Y Subscript s Baseline EndFraction StartFraction 4 bold-italic upper Y Subscript a comma 12 Superscript 2 Baseline bold-italic upper Y Subscript a comma s Superscript 2 Baseline Over StartAbsoluteValue def left-parenthesis bold-italic upper Y Subscript a Baseline right-parenthesis plus bold-italic upper Y Subscript a comma 11 Baseline bold-italic upper Y Subscript a comma r Baseline plus bold-italic upper Y Subscript a comma 22 Baseline bold-italic upper Y Subscript a comma s Baseline plus bold-italic upper Y Subscript a comma s Baseline bold-italic upper Y Subscript a comma r Baseline EndAbsoluteValue squared EndFraction 3rd Row 1st Column Blank 2nd Column equals StartFraction upper R e bold-italic upper Y Subscript r Baseline Over upper R e bold-italic upper Y Subscript s Baseline EndFraction StartFraction 4 Over StartAbsoluteValue StartFraction def left-parenthesis bold-italic upper Y Subscript a Baseline right-parenthesis Over bold-italic upper Y Subscript a comma 12 Baseline bold-italic upper Y Subscript a comma s Baseline EndFraction plus StartFraction bold-italic upper Y Subscript a comma 11 Baseline bold-italic upper Y Subscript a comma r Baseline Over bold-italic upper Y Subscript a comma 12 Baseline bold-italic upper Y Subscript a comma s Baseline EndFraction plus StartFraction bold-italic upper Y Subscript a comma 22 Baseline Over bold-italic upper Y Subscript a comma 12 Baseline EndFraction plus StartFraction bold-italic upper Y Subscript a comma r Baseline Over bold-italic upper Y Subscript a comma 12 Baseline EndFraction EndAbsoluteValue squared EndFraction EndLayout  (9.68)

and convert this into the transmission values using equation (9.12)

tau equals StartFraction upper R e bold-italic upper Y Subscript r Baseline Over upper R e bold-italic upper Y Subscript s Baseline EndFraction StartFraction 4 Over StartAbsoluteValue bold-italic upper T Subscript a comma 11 Baseline plus bold-italic upper T Subscript a comma 12 Baseline bold-italic upper Y Subscript a comma r Baseline plus StartFraction bold-italic upper T Subscript a comma 21 Baseline Over bold-italic upper Y Subscript a comma s Baseline EndFraction plus bold-italic upper T Subscript a comma 22 Baseline StartFraction bold-italic upper Y Subscript a comma r Baseline Over bold-italic upper Y Subscript a comma s Baseline EndFraction EndAbsoluteValue squared EndFraction  (9.69)

corresponding to expressions in the literature. When both media are the same, Ya,s=Ya,r, then

tau equals StartFraction 4 Over StartAbsoluteValue bold-italic upper T Subscript a comma 11 Baseline plus bold-italic upper T Subscript a comma 12 Baseline bold-italic upper Y Subscript a comma r Baseline plus StartFraction bold-italic upper T Subscript a comma 21 Baseline Over bold-italic upper Y Subscript a comma s Baseline EndFraction plus bold-italic upper T Subscript a comma 22 Baseline EndAbsoluteValue squared EndFraction  (9.70)

9.1.10.2 Absorption

The absorption is also derived in Chapter 2, namely by equation (2.105) for ϑ=0. The relevant quantity in this context is the input impedance. In the network context we have to solve the FE equation for loads at the input, say Q1. The input impedance follows from the resulting pressure p1

bold-italic upper Z Subscript a Baseline equals StartFraction bold-italic p 1 Over bold-italic upper Q 1 EndFraction  (9.71)

leading to the selected absorption depending on the impedance of the connected system.

alpha Subscript s Baseline equals left-parenthesis 1 minus StartAbsoluteValue bold-italic upper R EndAbsoluteValue squared right-parenthesis and bold-italic upper R equals StartFraction bold-italic upper Z Subscript a Baseline minus upper Z Subscript a comma s Baseline Over bold-italic upper Z Subscript a Baseline plus upper Z Subscript a comma s Baseline EndFraction  (9.72)

9.2 Coupled One-Dimensional Systems

In Section 9.1 the finite element representation of typical systems was developed to simulate acoustic networks. In the following we will elaborate some examples created by such subsystems.

9.2.1 Change in Cross Section

The changes in cross section motivated the selection of the volume flow as state variable. As a consequence the cross section element is supposed to be very simple, and the transfer matrix is the unit matrix

Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript change Baseline equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column 0 2nd Row 1st Column 0 2nd Column 1 EndMatrix  (9.73)

Assuming the same fluid on both sides, the transfer coefficient reads, with Ys=Ac(1)ρ0c0 and Yr=Ac(2)ρ0c0:

tau equals StartFraction upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline Over upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline EndFraction StartFraction 4 Over StartAbsoluteValue 1 plus StartFraction upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline Over upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline EndFraction EndAbsoluteValue squared EndFraction equals StartFraction 4 upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline Over left-parenthesis upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline plus upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline right-parenthesis squared EndFraction  (9.74)

9.2.2 Impedance Tube

The impedance tube is a device to measure the surface impedance of a specimen, for example a layer of foam, an absorber, or natural wall surfaces. In Figure 9.12 the impedance tube set-up is shown.

Figure 9.12 Impedance tube for measuring the surface impedance of a flat probe. Source: Alexander Peiffer.

We search for the dependency of the two measured pressures at positions x1 and x2 from the contact impedance at the end. In principle equation (4.15) can be applied directly, but in order to present the use of the transfer matrix method, we use it here.

The transfer matrices between x1 and x=0 as far as x1 and x2 are given by

Start 1 By 1 Matrix 1st Row upper T Subscript i Baseline EndMatrix equals Start 2 By 2 Matrix 1st Row 1st Column cosine left-parenthesis k upper L Subscript i Baseline right-parenthesis 2nd Column j StartFraction z 0 Over upper A Subscript c Baseline EndFraction sine left-parenthesis k upper L Subscript i Baseline right-parenthesis 2nd Row 1st Column j StartFraction upper A Subscript c Baseline Over z 0 EndFraction sine left-parenthesis k upper L Subscript i Baseline right-parenthesis 2nd Column cosine left-parenthesis k upper L Subscript i Baseline right-parenthesis EndMatrix with upper L 1 equals x 1 upper L 2 equals x 2 minus x 1  (9.75)

The end condition is given by (9.57) with Za=z/Ac. The pressure p1=p(x1) follows from

StartBinomialOrMatrix bold-italic p 1 Choose bold-italic upper Q 1 EndBinomialOrMatrix equals Start 1 By 1 Matrix 1st Row upper T 1 EndMatrix StartBinomialOrMatrix bold-italic upper Z slash upper A Subscript c Baseline Choose 1 EndBinomialOrMatrix  (9.76)

and p2=p(x2) from

StartBinomialOrMatrix bold-italic p 2 Choose bold-italic upper Q 2 EndBinomialOrMatrix equals Start 1 By 1 Matrix 1st Row upper T 2 EndMatrix StartBinomialOrMatrix bold-italic p 1 Choose bold-italic upper Q 1 EndBinomialOrMatrix  (9.77)

Using the pressure ratio H=p1/p2 and solving for Z gives:

bold-italic z equals j z 0 StartFraction bold-italic upper H sine left-parenthesis k x 2 right-parenthesis minus sine left-parenthesis k x 1 right-parenthesis Over bold-italic upper H cosine left-parenthesis k x 2 right-parenthesis minus cosine left-parenthesis k x 1 right-parenthesis EndFraction  (9.78)

9.2.3 Helmholtz Resonator

The Helmholtz resonator is the acoustic network representation of a tuned vibration absorber. Spring, mass, and damper are realised as fluid devices. The construction is as shown in Figure 9.13 on the left hand side.

Figure 9.13 Helmholtz (LHS) or quarter wave resonator (RHS). Source: Alexander Peiffer.

This cascade of involved subsystems can be readily described by transfer matrices. The neck of length L1 works as a mass of m=ρ0AcL1 and thus with transfer impedance Za,1D=jωρ0l1/Ac and resulting in the transfer matrix

Start 1 By 1 Matrix 1st Row upper T EndMatrix Subscript neck Baseline equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column j omega rho 0 StartFraction eff Over upper A Subscript c Baseline EndFraction 2nd Row 1st Column 0 2nd Column 1 EndMatrix  (9.79)

For the neck we follow the same argumentation as for the pores in the perforate absorber. We may have different situations as shown in figure 9.14. When there is no perforate at the opening, the resonator may use an end correction representing a fluid cylinder of length Lc on both sides if the jump in cross section is very high, thus Leff=L1+2Leff. As shown for the low frequency approximation for the piston in the wall (2.156), this length is Lc=0.85R with R being the radius.

Figure 9.14Neck tube with different end corrections. Source: Alexander Peiffer.

If the opening is covered by a perforate, this end correction can be replaced by the transfer impedance of the perforate. The correction length at the connection of the volume is taken into account by the length correction adjusting Leff. The end correction at the opening is included in the general transfer impedance term.

The end condition at the degree of freedom 2 is given by the volume form equation (9.22) and (9.57):

StartBinomialOrMatrix bold-italic p 2 Choose bold-italic upper Q 2 EndBinomialOrMatrix equals StartBinomialOrMatrix StartFraction rho 0 c 0 squared Over j omega upper V 0 EndFraction Choose 1 EndBinomialOrMatrix  (9.80)

The classical Helmholtz resonator is open, but in some applications the neck is covered by a perforate. In order to keep the option free, we represent the perforate by the generic transfer impedance Ra+jXa

Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript perf Baseline equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column upper R Subscript a Baseline plus j upper X Subscript a Baseline 2nd Row 1st Column 0 2nd Column 1 EndMatrix  (9.81)

In case of an open neck, the reactance is determined by the end correction 2Lc as discussed before:

upper X Subscript a Baseline equals omega rho 0 StartFraction upper L Subscript c Baseline Over upper A Subscript c Baseline EndFraction upper R Subscript a Baseline equals 0  (9.82)

When covered by a perforate, the transfer impedance is given by equation (9.50) converted into a radiation impedance

bold-italic upper Z Subscript a Baseline equals StartFraction bold-italic z Subscript perf Baseline Over upper A Subscript c Baseline EndFraction  (9.83)

The total transfer matrix follows from

Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript helm Baseline equals Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript perf Baseline Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript neck  (9.84)

and the surface impedance is given by

StartBinomialOrMatrix bold-italic p 1 Choose bold-italic upper Q 1 EndBinomialOrMatrix equals Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript perf Baseline Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript neck Baseline StartBinomialOrMatrix bold-italic upper Z Subscript a comma normal v normal o normal l normal u normal m normal e Baseline Choose 1 EndBinomialOrMatrix equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column j omega rho 0 StartFraction l 1 Over upper A Subscript c Baseline EndFraction plus left-parenthesis upper R plus j upper X right-parenthesis 2nd Row 1st Column 0 2nd Column 1 EndMatrix StartBinomialOrMatrix StartFraction rho 0 c 0 squared Over j omega upper V 0 EndFraction Choose 1 EndBinomialOrMatrix  (9.85)

Thus,

bold-italic upper Z Subscript a comma normal h normal e normal l normal m Baseline equals upper R Subscript a Baseline plus j left-parenthesis omega rho 0 StartFraction upper L 1 Over upper A Subscript c Baseline EndFraction plus upper X minus StartFraction rho 0 c 0 squared Over omega upper V 0 EndFraction right-parenthesis  (9.86)

The system is in resonance when the imaginary part is zero, and for the pure resonator without porous sheet we get

omega 0 equals StartRoot StartFraction c 0 squared Over upper V 0 EndFraction StartFraction upper A Subscript c Baseline Over upper L 1 plus 2 upper L Subscript c Baseline EndFraction EndRoot  (9.87)

At resonance the input impedance is Za,helm=Ra. Thus, for specific frequencies the Helmholtz resonator creates a matching end that would not be possible at low frequencies with such small dimensions. In Figure 9.15 the radiation impedance of one example resonator is shown. At resonance (ω0=1876.s1) the reactance curve crosses zero creating a purely resistive impedance at resonance frequency.

Figure 9.15 Radiation impedance of Helmholtz resonator of parameters R=2 mm, V0=1 cm3 and L1=5mm. Source: Alexander Peiffer.

When we need a perfectly matched end at a specific frequency, the Helmholtz resonator is the best choice for restricted space. Using the perforate of section 9.1.7.1 at the ending, the resonance moves to higher frequencies. This is because the upper mass is missing, but the resistance matches perfectly Za=ρ0c0/Ac. The results are shown in Section 9.2.4 in combination with the quarter wave absorber.

9.2.4 Quarter Wave Resonator

The quarter wave resonator is similar to the Helmholtz resonator except the fact that the resonator does not clearly separate between the mass (neck) and spring (volume) part. It consists of a tube with length l1 with rigid end and the same end correction as for the Helmholtz resonator neck. From section 4.1.1 and equation (4.17), we know the input impedance of the resonator.

bold-italic upper Z Subscript a comma normal p normal i normal p normal e Baseline equals StartFraction z 0 Over j upper A Subscript c Baseline tangent left-parenthesis bold-italic k upper L 1 right-parenthesis EndFraction with bold-italic k equals k left-parenthesis 1 minus j StartFraction eta Over 2 EndFraction right-parenthesis  (9.88)

The impedance is zero for L1=2n+14λ with n=0,1,2,3 so the resonance frequencies are

omega Subscript n Baseline equals StartFraction left-parenthesis 2 n plus 1 right-parenthesis pi c 0 Over l 1 EndFraction  (9.89)

With the transfer impedance for end correction of perforate consideration

StartBinomialOrMatrix bold-italic p 2 Choose bold-italic upper Q 2 EndBinomialOrMatrix equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column upper R Subscript a Baseline plus j upper X Subscript a Baseline 2nd Row 1st Column 0 2nd Column 1 EndMatrix StartBinomialOrMatrix bold-italic upper Z Subscript a comma normal p normal i normal p normal e Baseline Choose 1 EndBinomialOrMatrix  (9.90)

Deriving the input impedance from this we get

bold-italic upper Z Subscript a comma normal q normal u normal a normal r normal t normal e normal r Baseline equals StartFraction bold-italic p 2 Over bold-italic upper Q 2 EndFraction equals upper R Subscript a Baseline plus j upper X Subscript a Baseline plus StartFraction z 0 Over j upper A Subscript c Baseline tangent left-parenthesis bold-italic k upper L 1 right-parenthesis EndFraction  (9.91)

We may consider the transfer matrix of both end correction (9.82) and (9.83). Quarter wave resonators are rarely used without cover. They are covered in most cases by perforate, for example in the liner absorber of turbofan engines. The perforate from the last section can be used together with a resonator length of L2=10 mm.

In Figure 9.16 the impedances of both resonators are shown. The resonance frequency of the Helmholtz resonator is lower as for the quarter wave version. This results in better low frequency absorption as shown in Figure 9.17.

Figure 9.16 Radiation Impedance of Helmholtz resonator of parameters R=2 mm, V0=1 cm3, and l1=5 mm and quarter wave resonator l2=10 mm and perforated sheet. Source: Alexander Peiffer.

Figure 9.17Absorption coefficient of Helmholtz resonatorand quarter wave resonator. Source: Alexander Peiffer.

9.2.5 Muffler System

Mufflers are applied to reduce, for example, the noise of combustion engines or other machinery that creates pulsating volume flow in the audible frequency range. Here, we neglect any flow component by assuming that the flow speed is much lower than the speed of sound. More details for realistic mufflers can be found in Munjal (1987).

9.2.5.1 Expansion Chamber

A simple muffler consists of a combination of three tubes as shown in Figure 9.18. A pipe of cross section Ac(1) is expanded in the middle to a specific cross section Ac(2); this leads to 3 tube sections. As this is a cascaded set-up the transfer matrix method can be applied.

Figure 9.18 Expansion chamber arrangement a) reference system b) expansion system. Source: Alexander Peiffer.

The system is described by the matrix product of all sub pipes using equation (9.17) with according length and cross section

Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript EC Baseline equals Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript pipe 1 Baseline Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript pipe 2 Baseline Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript pipe 3  (9.92)

or we use the mobility form setting up the system matrix.

9.2.5.2 Open end Conditions

We start with the case that source impedance and end impedance are considered as one-dimensional free field Za,s=Zr,s=ρ0c0/Ac(1). In this situation the first and third pipe in Figure 9.18b can be neglected, and the total system performance is defined by the expanded chamber in the middle. From equation (9.70) we get

StartLayout 1st Row 1st Column tau 2nd Column equals StartFraction 4 Over StartAbsoluteValue cosine left-parenthesis bold-italic k upper L 2 right-parenthesis plus j StartFraction rho 0 c 0 Over upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline EndFraction StartFraction upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline Over rho 0 c 0 EndFraction sine left-parenthesis bold-italic k upper L 2 right-parenthesis plus j StartFraction rho 0 c 0 Over upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline EndFraction StartFraction upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline Over rho 0 c 0 EndFraction sine left-parenthesis bold-italic k upper L 2 right-parenthesis plus cosine left-parenthesis bold-italic k upper L 2 right-parenthesis EndAbsoluteValue squared EndFraction 2nd Row 1st Column Blank 2nd Column equals StartFraction 1 Over 1 plus one-fourth left-parenthesis StartFraction upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline Over upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline EndFraction plus StartFraction upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline Over upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline EndFraction right-parenthesis squared sine squared left-parenthesis bold-italic k upper L 2 right-parenthesis EndFraction EndLayout  (9.93)

or expressed as transmission loss

upper T upper L equals 10 log Subscript 10 Baseline left-parenthesis StartFraction 1 Over tau EndFraction right-parenthesis equals 10 log Subscript 10 Baseline left-bracket 1 plus one-fourth left-parenthesis StartFraction upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline Over upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline EndFraction plus StartFraction upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline Over upper A Subscript c Superscript left-parenthesis 2 right-parenthesis Baseline EndFraction right-parenthesis squared right-bracket  (9.94)

In Figure 9.19 different transmission losses are shown. The higher the change in cross section, the better the performance. We see that for sources in a well defined frequency range, a transmission loss of more than 30 dB can be achieved. However, when the center chamber is in resonance the transmission loss is zero, and the muffler does not work.

Figure 9.19Transmission loss of expansion chamber with various cross section ratios; l2=30 cm. Source: Alexander Peiffer.

9.2.5.3 Realistic End Conditions

Every muffler pipe radiates finally into the three-dimensional space. Thus, the pipe end will have the piston radiation impedance. In this case the first and third pipe of the system must be included, because reflections will occur at both ports.

The radiation impedance of the piston, according to equation (2.152), is:

bold-italic upper Z Subscript a comma r Baseline equals StartFraction rho 0 c 0 Over upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline EndFraction 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 with upper A Subscript c Superscript left-parenthesis 1 right-parenthesis Baseline equals pi upper R squared  (9.95)

When the matrix is built up as explained in section 9.1.2, we can consider each detail of the setup. We still assume an open condition at the source and compare the muffler to the reference pipe of same total length. The parameters of the expansion chamber are given in Table 9.1; the given center radius corresponds to an area ratio of 10.

Table 9.1 Expansion chamber geometry parameter

TubeLengthRadius
1L1=20 cm5 cm
2L2=30 cm15.8 cm
3L3=20 cm5 cm

The pressure magnitude at the end of the reference tube and expansion chamber is shown on figures 9.20 and 9.21. In the reference chamber the piston end condition leads to reflections that are seen at the entry. The expansion chamber shows several resonances coming from the the different pipe sections. Globally the pressure is reduced, but there are still some resonances where the muffler has a weak performance.

Figure 9.20 Pressure at entry (node1) and end (node4) of the reference tube with flanged end. Source: Alexander Peiffer.

Figure 9.21Pressure magnitude of expansion chamber at entry (node 1) and end (node 4), and at entry (node 2) and end (node 3) of the expansion chamber. Source: Alexander Peiffer.

The performance is qualified by the insertion loss comparing the pressure of reference and muffler system. The result is shown in figure 9.22. We see that at some frequencies, there are negative values, meaning that the reference system is more efficient than the muffler. Thus, real mufflers require additional damping, realized for example by steel wool, to take care of the resonances.

Figure 9.22 Insertion loss of expansion chamber. Source: Alexander Peiffer.

9.2.6 T-Joint

This system makes use of the Helmholtz resonator as a resonant damper in the one-dimensional propagation path. Such devices are called T-joints and are applied in hydraulic pipes to fight pulsation from hydraulic pumps or to reduce noise in the engine air intake. In Figure 9.23 a typical set-up is shown. The Helmholtz resonator is located in the middle, hence L1=L3=20 cm, the source impedance is open, and the end dynamics is given by the piston radiation impedance. The system can be described both ways: either the FE method using the Helmholtz resonator impedance as boundary condition at the center node, or by applying the transfer matrix method with a branch impedance.

Figure 9.23Pipe with t-joint and connected Helmholtz resonator. Source: Alexander Peiffer.

In this example we use Helmholtz resonators of the following parameters: RHR=1 cm, V0=100 cm3, and l2=2 cm. We use the pure Helmholtz resonator configuration and with a perforate cover in order to show the effect of damping. The perforate parameters are thickness h=0.2 mm, hole radius R=0.2 mm, and porosity σ=0.05. Such radiation impedance of both resonators is shown in Figure 9.24; we see a relative moderate resistivity of approximately 0.1, and the resonance of both is at ω0=2235 Hz.

Figure 9.24 Transfer impedance of T-joint perforate h=0.2 mm, R=0.2 mm, and σ=0.05. Source: Alexander Peiffer.

With the pure Helmholtz resonator, the insertion loss can be very high, as can be seen in Figure 9.25, but showing some negative loss at some resonances. The perforate version is not as effective but avoids the negative insertion loss.

Figure 9.25Insertion loss of T-joint system with and without perforate. Source: Alexander Peiffer.

9.2.7 Conclusions of 1D-Systems

In the last section a set of system and tools was developed to deal with one-dimensional fluid systems, namely pipes or tubes. Besides the practical use of such systems, the idea was to present means of noise control that are based on deterministic and coherent devices. From the point of view of the vibroacoustics engineer, the design and layout means adjusting the different resonances correctly.

9.3 Infinite Layers

Figure 9.26 Wavenumbers in x and z-directions for infinitely extended layers. Source: Alexander Peiffer.

Infinite layers are one-dimensional systems which are infinite in the other two dimensions. The infinity guaranties that plane waves can propagate in all directions, and the projection of the wave propagation normal to the plane is a one-dimensional wave motion. From the space vector definition in section 8.2.3, it is clear that the components kx,ky of the wavenumber k determines the in-plane state, whereas kz defines the propagation through the layer. For homogeneous layers we set ky=0 without loss of generality. We take an infinite layer of fluids and a plane wave impinging at angle θ; the wavenumber parallel to the surface is kx=ksinθ. The coordinate in the x-direction is given in wavenumber space by kx. In case of infinite layers – and only then – kx remains constant in each layer, and only kz varies. The in-depth wavenumber kz depends on the speed of sound in the layer. Thus, it is kz,i=ki2kx2. The transfer matrix method is converted to an infinite layer by adding a wavenumber argument to the transfer matrix.

StartLayout 1st Row 1st Column StartBinomialOrMatrix bold-italic p 1 left-parenthesis k Subscript x Baseline right-parenthesis Choose bold-italic v Subscript z Baseline Subscript 1 Baseline left-parenthesis k Subscript x Baseline right-parenthesis EndBinomialOrMatrix 2nd Column equals Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper T 11 left-parenthesis k Subscript x Baseline right-parenthesis 2nd Column bold-italic upper T 12 left-parenthesis k Subscript x Baseline right-parenthesis 2nd Row 1st Column bold-italic upper T 21 left-parenthesis k Subscript x Baseline right-parenthesis 2nd Column bold-italic upper T 22 left-parenthesis k Subscript x Baseline right-parenthesis EndMatrix StartBinomialOrMatrix bold-italic p 2 left-parenthesis k Subscript x Baseline right-parenthesis Choose bold-italic v Subscript z Baseline Subscript 1 Baseline left-parenthesis k Subscript x Baseline right-parenthesis EndBinomialOrMatrix EndLayout  (9.96)

9.3.1 Plate Layer

The mass layer can be enriched by bending stiffness effects using the results from section 8.2.4.3. Using the Ansatz of plane waves again this reads

normal upper Delta bold-italic p left-parenthesis x right-parenthesis equals normal upper Delta bold-italic p left-parenthesis k Subscript x Baseline right-parenthesis e Superscript minus j left-parenthesis k Super Subscript x Superscript right-parenthesis  (9.97)

For infinite plates the displacement must have the same wavenumber

bold-italic v Subscript z Baseline left-parenthesis x right-parenthesis equals j omega bold-italic w left-parenthesis k Subscript x Baseline x comma k Subscript y Baseline y right-parenthesis e Superscript minus j left-parenthesis k Super Subscript x Superscript x plus k Super Subscript y Superscript y right-parenthesis  (9.98)

entering this into the wave equation of the plate (3.206) leads to

upper B left-parenthesis k Subscript x Superscript 4 Baseline minus k Subscript upper B Superscript 4 Baseline right-parenthesis bold-italic w left-parenthesis k Subscript x Baseline comma k Subscript y Baseline right-parenthesis equals normal upper Delta bold-italic p left-parenthesis k Subscript x Baseline comma k Subscript y Baseline right-parenthesis  (9.99)

and the transfer impedance is

bold-italic z left-parenthesis k Subscript x Baseline right-parenthesis equals StartFraction normal upper Delta bold-italic p left-parenthesis k Subscript x Baseline right-parenthesis Over bold-italic v Subscript z Baseline left-parenthesis k Subscript x Baseline right-parenthesis EndFraction equals StartFraction normal upper Delta bold-italic p left-parenthesis k Subscript x Baseline right-parenthesis Over j omega bold-italic w left-parenthesis k Subscript x Baseline right-parenthesis EndFraction equals j omega upper B left-bracket left-parenthesis k Subscript x Baseline x squared plus k Subscript y Baseline y squared right-parenthesis squared minus k Subscript upper B Superscript 4 Baseline right-bracket  (9.100)

After some modification the transfer impedance reads as:

bold-italic z left-parenthesis k Subscript x Baseline right-parenthesis equals j omega m double-prime left-parenthesis 1 minus StartFraction k Subscript x Superscript 4 Baseline Over bold-italic k Subscript upper B Superscript 4 Baseline EndFraction right-parenthesis with bold-italic k Subscript upper B Baseline left-parenthesis omega right-parenthesis equals StartFraction m double-prime omega squared Over upper B EndFraction  (9.101)

For zero bending stiffness B0, the transfer impedance is Z(kx)=jωm showing that the limp behavior does not depend on the wavenumber kx. As there is no in-depth wave propagation, the transfer matrix has the same form as (9.27) with the above given transfer impedance.

StartLayout 1st Row 1st Column Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript plate 2nd Column equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column bold-italic z left-parenthesis k Subscript x Baseline right-parenthesis 2nd Row 1st Column 0 2nd Column 1 EndMatrix EndLayout  (9.102)

9.3.2 Lumped Elements Layers

In general, the parameters of lumped elements don’t change due to the wavenumber argument. So, mass layers and perforates can be used as is. The given transfer impedances according to (9.27) or specifically for the mass layer (9.29) or the perforate (9.50) are still valid.

StartLayout 1st Row 1st Column Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript lumped 2nd Column equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column bold-italic z Subscript 1 upper D Baseline 2nd Row 1st Column 0 2nd Column 1 EndMatrix EndLayout 

9.3.3 Fluid Layer

The transfer matrix method of one layer excited by a plane wave of kx,kz follows from the solutions of the wave equation in the z-direction. We slightly modify equation (4.1)

StartLayout 1st Row 1st Column bold-italic p Subscript i Baseline left-parenthesis z comma k Subscript x Baseline right-parenthesis equals 2nd Column bold-italic upper A e Superscript minus j bold-italic k Super Subscript z comma i Superscript z plus bold-italic upper B e Superscript j bold-italic k Super Subscript z comma i Superscript z EndLayout  (9.103)
StartLayout 1st Row 1st Column bold-italic v Subscript z Baseline Subscript i Baseline left-parenthesis z comma k Subscript x Baseline right-parenthesis equals minus StartFraction 1 Over j omega bold-italic rho Subscript i Baseline EndFraction StartFraction partial-differential bold-italic p Subscript i Baseline Over partial-differential z EndFraction equals StartFraction bold-italic k Subscript z comma i Baseline Over omega bold-italic rho Subscript i Baseline EndFraction left-parenthesis 2nd Column bold-italic upper A e Superscript minus j bold-italic k Super Subscript z comma i Superscript z Baseline minus bold-italic upper B e Superscript j bold-italic k Super Subscript z comma i Superscript z Baseline right-parenthesis EndLayout  (9.104)

keeping in mind that

bold-italic k Subscript z comma i Baseline left-parenthesis k Subscript x Baseline right-parenthesis equals StartRoot bold-italic k Subscript i Superscript 2 Baseline minus k Subscript x Superscript 2 Baseline EndRoot  (9.105)

is a function of kx. We get in accordance with the one-dimensional fluid transfer matrix method

StartLayout 1st Row 1st Column Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript i comma normal f normal l normal u normal i normal d Baseline left-parenthesis k Subscript x Baseline right-parenthesis 2nd Column equals Start 2 By 2 Matrix 1st Row 1st Column cosine left-parenthesis bold-italic k Subscript z comma i Baseline h right-parenthesis 2nd Column j StartFraction omega bold-italic rho Subscript i Baseline Over bold-italic k Subscript z comma i Baseline EndFraction sine left-parenthesis bold-italic k Subscript z comma i Baseline h right-parenthesis 2nd Row 1st Column j StartFraction bold-italic k Subscript z comma i Baseline Over omega bold-italic rho Subscript i Baseline EndFraction sine left-parenthesis bold-italic k Subscript z comma i Baseline h right-parenthesis 2nd Column cosine left-parenthesis bold-italic k Subscript z comma i Baseline h right-parenthesis EndMatrix EndLayout  (9.106)

9.3.4 Equivalent Fluid – Fiber Material

Fiber materials are used for acoustic isolation and absorption. The acoustic fluid motion in the fiber network leads to losses in the flow and wave propagation. The deceleration of the fluid motion acts on the fiber matrix and accelerates the fibers. A model for such porous materials is mandatory for noise control application, but the required theory is out of scope for this book. Allard et al. (2005) provides a comprising overview about the acoustic theory and application of porous material.

Figure 9.27 Sketch of fiber absorber. Source: Alexander Peiffer.

We use the model of the limb equivalent fluid that was developed by Champoux for the rigid or fixed matrix Champoux and Stinson (1992) and extended by the limb frame model as described, for example, by Paneton Panneton (2007). The useful thing with this model is that the acoustics in the fiber material are still described by acoustic fluid parameters. They become complex and frequency dependent, but the existing models of fluid layers or tubes can still be used. A thorough description by complex models – as for example Biot models Biot (1962) – considering the two coupled waves in the solid and fluid phase, require a completely different approach handling several degrees of freedom and connectivity conditions. In order to keep the efforts reasonable, we leave it with the simple presentation of the final formulas:

StartLayout 1st Row 1st Column upper G Subscript j Baseline left-parenthesis omega right-parenthesis 2nd Column equals left-parenthesis 1 plus StartFraction 4 j alpha Subscript infinity Superscript 2 Baseline eta Subscript upper V Baseline rho 0 omega Over sigma squared upper Lamda squared upper Phi squared EndFraction right-parenthesis Superscript 1 slash 2 EndLayout  (9.107a)
StartLayout 1st Row 1st Column rho Subscript eq 2nd Column equals StartFraction alpha Subscript infinity Baseline rho 0 Over upper Phi EndFraction left-parenthesis 1 plus StartFraction sigma upper Phi Over j omega rho 0 alpha Subscript infinity Baseline EndFraction upper G Subscript j Baseline left-parenthesis omega right-parenthesis right-parenthesis EndLayout  (9.107b)
StartLayout 1st Row 1st Column rho Subscript limp 2nd Column equals StartFraction rho Subscript eq Baseline rho Subscript bulk Baseline minus rho 0 squared Over rho Subscript bulk Baseline plus rho Subscript e q Baseline minus 2 rho 0 EndFraction EndLayout  (9.107c)
StartLayout 1st Row 1st Column upper K 2nd Column equals StartFraction rho 0 c 0 squared Over kappa minus left-parenthesis kappa minus 1 right-parenthesis left-bracket 1 plus StartFraction 8 eta Over j upper Lamda prime squared upper P r omega rho 0 EndFraction StartRoot 1 plus j rho 0 StartFraction omega upper P r upper Lamda prime squared Over 16 eta EndFraction EndRoot right-bracket Superscript negative one-half Baseline EndFraction EndLayout  (9.107d)
StartLayout 1st Row 1st Column z 2nd Column equals StartRoot upper K rho Subscript limp Baseline EndRoot EndLayout  (9.107e)
StartLayout 1st Row 1st Column upper Gamma 2nd Column equals j omega StartRoot rho Subscript limp Baseline slash upper K EndRoot EndLayout  (9.107f)
StartLayout 1st Row 1st Column c 2nd Column equals j omega slash upper Gamma EndLayout  (9.107g)

There is a confusing variety of parameters that are not all independent as shown by Horoshenkov et al. (2019). A heuristic explanation of the parameters is given in tables 9.2 and 9.3.

Table 9.2 Fiber parameters of the equivalent fluid model

SymbolDescription
Φ=VfiberVallVolume porosity, fraction of fiber and total volume
αTortuosity, ratio of average path through the absorber to straight path,
a measure for the diversion of the fluid
ΛViscous characteristic length
ΛThermal characteristic length
σStatic air flow resistivity
ρbulkApparent total density of fluid and fiber
ρeqEquivalent density of the fluid in a rigid/fixed fiber matrix
ρlimpEquivalent density of the limp equivalent fluid

Table 9.3Fluid parameters

SymbolDescription
ρ0Fluid density
ηVDynamic viscosity
PrPrandtl number

Table 9.4 Material parameters of soft fiber material. Source: Panneton (2007).

SymbolValueUnits
Φ0.98
σflow25 000N s/m4
α1.02
Λ190μm
ρ01.208kg/m3
ρbulk31.1kg/m3
1 For fibrous material Λ=2Λ can be assumed.

The soft fibrous material from Panneton (2007) is taken as an example. In figures 9.29 and 9.28 the results for the complex speed of sound and the density show a strong dependency on frequency. The are two major versions of equivalent fluid models: the rigid frame model, assuming a fixed fiber matrix with no participation of the fibers to the motion, and the limp frame model that considers forced motion of the frame inertia. One can easily switch from the rigid to the limp model by using equation (9.107)c. For low frequencies the real density of the limp model reaches values near the bulk density of the material, because the material moves with the inertia of the bulk material. The rigid frame model does not catch this effect. In the high frequency limit, both models provide the same result. According to Panneton the high frequency limit is

limit Underscript omega right-arrow normal infinity Endscripts rho Subscript e q Baseline equals StartFraction alpha Subscript normal infinity Baseline rho 0 Over normal upper Phi EndFraction  (9.108)

Figure 9.28 Dynamic complex density of limp and rigid fiber material. Source: Alexander Peiffer.

Figure 9.29Dynamic complex sound speed of limp and rigid fiber material. Source: Alexander Peiffer.

A similar effect can be seen for the speed of sound. For high frequencies the real value of the limp model reaches the real value of the rigid model. At low frequencies the speed of sound of the fiber material is below the fluid sound speed because of the additional density from the bulk material. In Panneton (2007) the results are compared to tests, showing that the rigid frame model fails for low frequencies.

To conclude, for the calculation of fiber material sound propagation, we presented a useful model that is, for example, used frequently in aerospace applications (Moeser et al., 2008).

9.3.5 Performance Indicators

The performance indicators from the last section are reused here but with wavenumber or angle as argument and now formulated for the impedance instead of radiation mobility. The source and receiver impedances of the fluids are also dependant on the angle of attack due to equation (2.116). With zs/rzs/r/cos(ϑ) we get

tau left-parenthesis theta right-parenthesis equals StartFraction upper R e bold-italic z Subscript s Baseline Over upper R e bold-italic z Subscript r Baseline EndFraction StartFraction 4 Over StartAbsoluteValue bold-italic upper T 11 left-parenthesis k Subscript x Baseline right-parenthesis plus StartFraction bold-italic upper T 12 left-parenthesis k Subscript x Baseline right-parenthesis cosine left-parenthesis theta right-parenthesis Over bold-italic z Subscript r Baseline EndFraction plus StartFraction bold-italic upper T 21 left-parenthesis k Subscript x Baseline right-parenthesis bold-italic z Subscript s Baseline Over cosine left-parenthesis theta right-parenthesis EndFraction plus bold-italic upper T 22 left-parenthesis k Subscript x Baseline right-parenthesis StartFraction bold-italic z Subscript r Baseline Over bold-italic z Subscript s Baseline EndFraction EndAbsoluteValue squared EndFraction  (9.109)

The diffuse field performance is derived by equations (8.13) or (8.14). The above formula is often used with an empirical maximum angle ϑmax leading to the formula shown in (8.103). When we are interested in the absorption, we have similar modifications for (9.72)

alpha Subscript s Baseline left-parenthesis theta right-parenthesis equals left-parenthesis 1 minus StartAbsoluteValue bold-italic upper R left-parenthesis theta right-parenthesis EndAbsoluteValue squared right-parenthesis and bold-italic upper R left-parenthesis theta right-parenthesis equals StartStartFraction bold-italic z Subscript s Baseline left-parenthesis theta right-parenthesis minus StartFraction z 0 Over cosine left-parenthesis theta right-parenthesis EndFraction OverOver bold-italic z Subscript s Baseline left-parenthesis theta right-parenthesis plus StartFraction z 0 Over cosine left-parenthesis theta right-parenthesis EndFraction EndEndFraction  (9.110)

getting an according diffuse field absorption by angular averaging

left pointing angle alpha Subscript s Baseline right pointing angle equals 2 integral Subscript zero width space zero width space zero width space 0 Superscript zero width space zero width space zero width space pi slash 2 Baseline alpha Subscript s Baseline left-parenthesis theta right-parenthesis cosine left-parenthesis theta right-parenthesis sine left-parenthesis theta right-parenthesis d theta  (6.72)

For normal irradiation the perfect absorption is achieved for zs=ρ0c0, thus a perfectly matching impedance with zero reactance and resistance equal to the characteristic impedance of air. However, this condition cannot be met for all angles of incidence. In the diffuse field integration the most important angle is ϑ=45, because the sinϑcosϑ-term in equation (6.70) has its maximum at this angle.

In Figure 9.30 the diffuse field absorption is calculated for different values of the real surface impedance. We see that highest diffuse field absorption can be achieved by zs=az0 with a[1,1.5]. So, the design goal for a best diffuse field absorbing device is given by this rule.

Figure 9.30Diffuse field absorption for varying values of zs. Source: Alexander Peiffer.

9.3.6 Conclusions on Layer Formulation

The above derived transfer matrices are useful for the determination of transmission (and therefore also coupling) of large area junctions. Energy is removed from the reverberant field by absorption taking place at every reflection. The approximation of infinite layer is the more valid the larger the areas are. In any case this formulation is a powerful tool to derive properties of so-called acoustic treatments, layers of different materials that are used as noise control treatment in many applications.

9.4 Acoustic Absorber

Acoustic absorbers are used to reduce the noise levels in cavities and rooms. The task is to create a lay-up that maximizes the absorption in a specified frequency range with the least space and weight requirements. The target value is zs=2ρ0c0, because this leads to perfect matching at an angle of 45 that contributes most to the diffuse absorption coefficient. We start with a simple absorber created by a single layer of fiber material. The next step will be a combined absorber consisting of a fiber absorber plus perforated sheet.

9.4.1 Single Fiber Layer

When placing an absorbing fiber layer in front of a rigid wall, the transfer matrix is connected to a rigid wall, meaning that v2=0 in equation (9.96)

StartBinomialOrMatrix bold-italic p 1 left-parenthesis k Subscript x Baseline right-parenthesis Choose bold-italic v Subscript z Baseline Subscript 1 Baseline left-parenthesis k Subscript x Baseline right-parenthesis EndBinomialOrMatrix equals Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper T 11 left-parenthesis k Subscript x Baseline right-parenthesis 2nd Column bold-italic upper T 12 left-parenthesis k Subscript x Baseline right-parenthesis 2nd Row 1st Column bold-italic upper T 21 left-parenthesis k Subscript x Baseline right-parenthesis 2nd Column bold-italic upper T 22 left-parenthesis k Subscript x Baseline right-parenthesis EndMatrix StartBinomialOrMatrix bold-italic p 2 left-parenthesis k Subscript x Baseline right-parenthesis Choose 0 EndBinomialOrMatrix equals StartBinomialOrMatrix bold-italic upper T 11 left-parenthesis k Subscript x Baseline right-parenthesis Choose bold-italic upper T 21 left-parenthesis k Subscript x Baseline right-parenthesis EndBinomialOrMatrix bold-italic p 2 left-parenthesis k Subscript x Baseline right-parenthesis  (9.111)

For the surface impedance this leads to

bold-italic z left-parenthesis k Subscript x Baseline right-parenthesis equals StartFraction bold-italic p 1 left-parenthesis k Subscript x Baseline right-parenthesis Over bold-italic v Subscript z Baseline left-parenthesis k Subscript x Baseline right-parenthesis EndFraction equals StartFraction bold-italic upper T 11 left-parenthesis k Subscript x Baseline right-parenthesis Over bold-italic upper T 21 left-parenthesis k Subscript x Baseline right-parenthesis EndFraction equals StartFraction bold-italic z 0 Over j tangent left-parenthesis bold-italic k h right-parenthesis EndFraction  (9.112)

For first evaluation of our fiber material described in section 9.3.4, we use the fluid transfer matrix (9.106) with the material parameters from (9.107a)–(9.107)g. In figure 9.31 the impedance spectrum for ϑ=0 is shown for materials of different thicknesses.

Figure 9.31 Perpendicular surface impedance of single fiber layer in front of a rigid wall. Source: Alexander Peiffer.

The reactance of the 20 cm layer shows a first λ/2 resonance at ω=650Hz but with no clear peak. The 10 cm version has a local maximum around the double frequency. As damping increases with frequency, the 10 cm resonance is weaker and does not reach zero reactance. When we consider equation (2.104), at higher frequency both layers coincide, because the high damping in the material prevents any feedback from reflected waves.

We see that zero reflection requires a real zs=zs=ρ0c0 for normal incidence. So, the material shown before seems to be a performing fiber absorber, as resistance is near the impedance of air. In Figure 9.32 the absorption due to normal wave incidence is shown for both layers. The early resonance of the 20 cm layer leads to better performance at lower frequencies with a dip afterwards. At ω = 400s−1, there is weak absorption for the 10 cm layer, whereas the 20 cm layer provides already α0.75. However, a 20 cm layer requires a lot of space that is not always available. Thus, we look for an option to reduce the thickness but keep the low frequency performance.

Figure 9.32Perpendicular surface absorption of a single fiber layer in front of a rigid wall. Source: Alexander Peiffer.

The low speed of sound in the fiber material leads to a large wavenumber in the fiber. The consequence is that even oblique waves are diffracted to the surface normal, and the angle dependence to the impedance is quite low. Thus, the normal and diffuse absorption are not that different (figure 9.33).

Figure 9.33 Diffuse absorption of a single layer in front of a rigid wall. Source: Alexander Peiffer.

9.4.2 Multiple Layer Absorbers

In many cases there is not enough space for large thickness absorbers available; or, the environment does not allow for porous materials, because the surface will be exposed to dirt and humidity. we must protect the absorber by a thin layer (thin and soft plates) or use perforate that can be cleaned.

9.4.3 Absorber with Perforate

A perforate with micro absorption may have a transfer impedance with resistance on the order of magnitude of air and with additional mass generated by the neck effect as described in section 9.1.7. This mass can be used to lower the resonance of the total system and therefore to reduce the thickness of the absorber.

The advantage of this concept is that we can separately tune the resistance of the front sheet to air and the thickness of fluid or fiber layer to create zero reactance. Let us assume that the layer generates a surface impedance of zs2. The state vector is then given by

StartBinomialOrMatrix bold-italic p 2 Choose bold-italic v 2 EndBinomialOrMatrix equals StartBinomialOrMatrix bold-italic z Subscript s Baseline 2 Baseline Choose 1 EndBinomialOrMatrix bold-italic v 2  (9.113)

The state vector of the front layer follows from matrix multiplication with the limp layer transfer matrix

StartBinomialOrMatrix bold-italic p 1 Choose bold-italic v 1 EndBinomialOrMatrix equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column bold-italic z Subscript perf Baseline 2nd Row 1st Column 0 2nd Column 1 EndMatrix StartBinomialOrMatrix bold-italic z Subscript s Baseline 2 Baseline Choose 1 EndBinomialOrMatrix bold-italic v 2 equals StartBinomialOrMatrix bold-italic z Subscript perf Baseline plus bold-italic z Subscript s Baseline 2 Baseline Choose 1 EndBinomialOrMatrix bold-italic v 2  (9.114)

so the final surface impedance reads

bold-italic z Subscript s Baseline 1 Baseline equals StartFraction bold-italic p 1 Over bold-italic v 1 EndFraction equals bold-italic z Subscript perf Baseline plus bold-italic z Subscript s Baseline 2  (9.115)

This is why limb elements that are characterised by v1=v2 are well represented by a transfer impedance that simply adds to the backing impedance. Coming back to our design problem, we can select an air layer that provides zero reactance at thickness h=nλ/2 and choose a perforate with the matching resistivity. In Fuchs and Zha (1995) some micro-perforate absorbers with the desired quantities are given. We use the absorber of figure 6 of Fuchs and Zha and the parameters shown in Table 9.5.

Table 9.5 Absorbers with different perforate plates and constant surface porosity. Source: Fuchs and Zha (1995).

Mesht/mmR/mmd/mmh/mmσ
13.01.522.5500.014 0
23.00.2253.37500.014 0
33.00.0751.13500.013 8

In figures 9.34 and 9.35 the resistance and reactance of the perforates are shown. One can recognize that the resistance of the second mesh fits best to our above condition for best absorption. The resistance of mesh 1 is too low and mesh 3 too high. Mesh 2 nearly meets the requirements over a large frequency range.

Figure 9.34Transfer resistance of three different micro-perforates. Source: Alexander Peiffer.

Figure 9.35 Transfer reactance of three different micro-perforates. Source: Alexander Peiffer.

From the three reactance curves, we conclude that the high inertia from mesh 1 and 2 might reduce the frequency range of the absorber, because the reactance of the mesh is added to the reactance of the air spring as shown in Figure 9.35. A pure resistive absorber would have best absorption for X=0; with the mass effect, the resonance is lower. This is because the stiffness of the air spring sees the mass of the perforate, and the resonance frequency is reduced. In Figure 9.35 this can be seen by the intersection point of X perf and Xair.

The normal impedance of the full absorber is shown in Figure 9.36. The normal and diffuse field absorption is shown in figures 9.37 and 9.38. The result does not fit perfectly to Fuchs’ test results, but the global tendency and resonance frequency is well met.

Figure 9.36Surface impedance of micro-perforate absorber. Source: Alexander Peiffer.

Figure 9.37 Normal incidence absorption of micro-perforate absorber. Source: Alexander Peiffer.

Figure 9.38Diffuse field absorption of micro-perforate absorber. Source: Alexander Peiffer.

9.4.4 Single Degree of Freedom Liner

The micro-perforate absorber is a perfect candidate for building absorbers. In aerospace applications the absorber must withstand strong aerodynamic load, pressure, and heat. Thus, the perforate is fixed to a carrier structure. This is usually a honeycomb material made out of aluminium or aramid fiber paper (Figure 9.39). Such liners are used in the inlet section of turbo-fan nacelles but also inside engines in cold and hot streams. In these absorbers the waves cannot propagate in in-plane directions, and we must use the normal incidence impedance of an air layer as surface impedance of the honeycomb layer. This means we set the wave number in the x-direction kx=0 in equation (9.105) and neglect the volume fraction of the honeycomb wall material.

Figure 9.39Single degree of freedom liner with (a) hard backing, (b) honeycomb core and (c) perforated sheet. Source: Alexander Peiffer.

With the configuration from Table 9.6, we get the transfer impedances as shown in figure 9.40. We see that the resistance condition R=z0 and the λ/2 resonance occur at the same frequency. In Figure 9.41 we see that the absorption is nearly 1 at a certain frequency band around this regime. This is acceptable, because in turbofan engines the absorption is optimized for specific tones in the engine. It must be noted that in real aircraft liner designs, the airflow must be considered, because the high Mach number current changes the neck correction (Hubbard and Acoustical Society of America, 1995).

Figure 9.40 Transfer resistance and reactance of liner perforate. Source: Alexander Peiffer.

Figure 9.41Surface impedance and absorption of liner. Source: Alexander Peiffer.

Table 9.6 Aerospace liner set-up

t/mmR/mmd/mmh/mmσ
1.50.23300.022

9.5 Acoustic Wall Constructions

9.5.1 Double Walls

Inspecting wall configurations that aim at high acoustic isolation at reasonable weight leads to the conclusion that many engineering systems are enhancing the acoustic isolation of a single wall by simply adding a second wall with some absorbing material in the middle. Examples are the interior lining in aircraft covering layers of glass fiber blankets (Peiffer et al., 2007, 2013), gypsum plasterboards in buildings, or mass-spring layers in the automotive industry used, for example, to increase the isolation of the firewall in cars.

The generic setup is shown in Figure 9.42. The walls consist of panels that may be considered as limp mass or plates. In both cases they are described by thin layer transfer equation (9.27) with different transfer impedances z1 and z2

StartLayout 1st Row 1st Column Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript 1 2nd Column equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column bold-italic z 1 2nd Row 1st Column 0 2nd Column 1 EndMatrix 3rd Column Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript 2 4th Column equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column bold-italic z 2 2nd Row 1st Column 0 2nd Column 1 EndMatrix EndLayout  (9.116)

Figure 9.42 Double wall lay-up of panel–cavity–panel. Source: Alexander Peiffer.

The cavity or fluid layer is given by (9.106).

The transfer matrix of the full system is given by

Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript upper D upper W Baseline equals Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript 1 Baseline Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript fluid Baseline Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript 2  (9.117)

In this case the result can be derived analytically and we get after some math

StartLayout 1st Row 1st Column Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript upper D upper W 2nd Column equals Start 2 By 2 Matrix 1st Row 1st Column bold-italic upper T 11 2nd Column bold-italic upper T 12 2nd Row 1st Column bold-italic upper T 21 2nd Column bold-italic upper T 22 EndMatrix 2nd Row 1st Column bold-italic upper T 11 2nd Column equals StartFraction j bold-italic z 1 bold-italic k Subscript z Baseline sine left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis Over omega bold-italic rho EndFraction plus cosine left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis 3rd Row 1st Column bold-italic upper T 12 2nd Column equals left-parenthesis bold-italic z 1 plus bold-italic z 2 right-parenthesis cosine left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis plus j left-parenthesis StartFraction bold-italic z 1 bold-italic z 2 bold-italic k Subscript z Baseline Over omega bold-italic rho EndFraction plus StartFraction omega bold-italic rho Over bold-italic k Subscript z Baseline EndFraction right-parenthesis sine left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis 4th Row 1st Column bold-italic upper T 21 2nd Column equals StartFraction j bold-italic k Subscript z Baseline sine left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis Over omega bold-italic rho EndFraction 5th Row 1st Column bold-italic upper T 22 2nd Column equals StartFraction j bold-italic z 2 bold-italic k Subscript z Baseline sine left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis Over omega bold-italic rho EndFraction plus cosine left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis EndLayout  (9.118)

Applying equation (9.109) gives the transmission coefficient for the double leaf configuration

StartLayout 1st Row 1st Column tau 2nd Column equals StartAbsoluteValue ContinuedFraction minus 2 j cosine left-parenthesis theta right-parenthesis slash cosine left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis Over StartLayout 1st Row StartFraction bold-italic k Subscript z Baseline Over omega bold-italic rho EndFraction left-parenthesis z 0 plus bold-italic z 1 plus bold-italic z 2 right-parenthesis cosine left-parenthesis theta right-parenthesis tangent left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis minus 2 j cosine left-parenthesis theta right-parenthesis midline-horizontal-ellipsis 2nd Row midline-horizontal-ellipsis plus left-bracket left-parenthesis StartFraction bold-italic z 1 bold-italic z 2 bold-italic k Subscript z Baseline Over omega bold-italic rho z 0 EndFraction plus StartFraction omega bold-italic rho Over bold-italic k Subscript z Baseline z 0 EndFraction right-parenthesis tangent left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis minus j StartFraction bold-italic z 1 plus bold-italic z 2 Over z 0 EndFraction right-bracket cosine squared left-parenthesis theta right-parenthesis EndLayout EndAbsoluteValue squared EndLayout  (9.119)

For low frequencies |kzh|1 the wavelength in the double wall cavity is much larger than the thickness, and equation (9.118) can be simplified to

limit Underscript k Subscript z Baseline h right-arrow 0 Endscripts Start 1 By 1 Matrix 1st Row bold-italic upper T EndMatrix Subscript upper D upper W Baseline equals Start 2 By 2 Matrix 1st Row 1st Column 1 2nd Column bold-italic z 1 plus bold-italic z 2 2nd Row 1st Column 0 2nd Column 1 EndMatrix  (9.120)

Thus, at low frequencies the layers are supposed to be stiffly connected.

9.5.1.1 Double Wall of Limp Mass

For further insight we use the ideal case of two limp mass layers with transfer impedance z1=jm1ω and z2=jm2ω. The fluid gap can be a fluid like air, other gases, or an equivalent fluid, hence kz=kcos(ϑ). The double wall transmission loss reads with this assumption

StartLayout 1st Row 1st Column tau 2nd Column equals StartAbsoluteValue ContinuedFraction minus 2 j cosine left-parenthesis theta right-parenthesis slash cosine left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis Over StartLayout 1st Row StartFraction bold-italic k Subscript z Baseline Over omega bold-italic rho EndFraction left-parenthesis z 0 plus j left-parenthesis m double-prime 1 plus m double-prime 2 right-parenthesis omega right-parenthesis cosine left-parenthesis theta right-parenthesis tangent left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis minus 2 j cosine left-parenthesis theta right-parenthesis midline-horizontal-ellipsis 2nd Row midline-horizontal-ellipsis plus left-bracket left-parenthesis minus StartFraction m double-prime 1 m double-prime 2 omega bold-italic k Subscript z Baseline Over bold-italic rho z 0 EndFraction plus StartFraction omega bold-italic rho Over bold-italic k Subscript z Baseline z 0 EndFraction right-parenthesis tangent left-parenthesis h bold-italic k Subscript z Baseline right-parenthesis plus StartFraction left-parenthesis m double-prime 1 plus m double-prime 2 right-parenthesis omega Over z 0 EndFraction right-bracket cosine squared left-parenthesis theta right-parenthesis EndLayout EndAbsoluteValue squared EndLayout  (9.121)

We can show that at low frequencies |kzh1|, eq. (9.121) gives the mass law of infinite single walls (8.96) with total mass mt=m1+m2. The same conclusion follows from (9.120) when we assume the mass transfer impedance z=jmtω. In figure 9.43 the transmission coefficients are shown for ϑ=0 and ϑ=50. There is a first minimum of the transmission loss that is called the double wall resonance and will be dealt with later. Below this resonance the result coincides with the mass law curve of a single wall of the same total mass.

Figure 9.43Transmission coefficient for plane waves transmission. Double wall of two limp layers of m1=m2=1 kg/m2 and air cavity of h=10 cm thickness. At the axis the resonance frequencies of perpendicular waves are denoted. Source: Alexander Peiffer.

In the regime above the double wall resonance, there are several dips that correspond to resonances of the cavity for kzh=nπ, and the tangent in equation (9.121) becomes infinite. In other words the thickness of the cavity equals integer multiples of the half wavelength. The cavity resonances are denoted by ωi in figure 9.43. When damping in the cavity is not too high (which is definitely the case for air), the anti-resonances for

kzh=π(2n1)/2 lead to a decoupling and equation (9.121) is approximated for high frequencies by:

tau equals StartAbsoluteValue StartFraction j Baseline 2 omega bold-italic rho z 0 Over omega squared m double-prime 1 m double-prime 2 cosine squared theta bold-italic k EndFraction EndAbsoluteValue squared equals left-parenthesis StartFraction 2 z 0 Over omega m double-prime 1 cosine theta EndFraction right-parenthesis squared left-parenthesis StartFraction 2 z 0 Over omega m double-prime 2 cosine theta EndFraction right-parenthesis squared left-parenthesis StartFraction bold-italic z Over 2 z 0 EndFraction right-parenthesis squared  (9.122)

Note that high damping means a complex wavenumber and the condition kzh=πsin((2n1)/2)1 cannot be met, and the above estimation is not valid. Writing the above expression as transmission loss, we see that above the double wall resonance, the single transmission loss of each wall is summed up, plus an extra term

upper T upper L equals upper T upper L left-parenthesis m double-prime 1 right-parenthesis plus upper T upper L left-parenthesis m double-prime 2 right-parenthesis plus 6 dB plus 10 log Subscript 10 Baseline StartAbsoluteValue StartFraction z 0 Over bold-italic z EndFraction EndAbsoluteValue  (9.123)

The best use of available mass for high transmission loss is given for a distribution of two walls. This is the reason why double leaf constructions are the power weapons for acoustic isolation in many fields of technical acoustics.

However, at the resonances the isolation performance of the double wall can be low, lower than the mass law. Thus, the resonances must be reduced by damping in the cavity, and we must understand what determines the frequency of the double wall resonance. This can be found by determining the minimum of the denominator of equation (9.121) as shown, for example, by Fahy (1985). We apply the transfer matrix expression (9.118). Perfect transmission is given when the transfer matrix becomes a unit-matrix, thus T11=T22=1 and T12=T21=0. In general the thickness of a double wall is much smaller than the wavelength at low frequencies, so hkz1 and we use sin(kzh)kzh and cos(kzh)1.

The diagonal components of (9.118) are, with zii=jmiω

bold-italic upper T Subscript i i Baseline almost-equals 1 minus StartFraction m double-prime Subscript i Baseline omega squared h Over bold-italic rho bold-italic c squared cosine squared theta EndFraction  (9.124)

and are close to unity for typical wall dimensions. The approximation of T12 reads

StartLayout 1st Row 1st Column bold-italic upper T 12 2nd Column almost-equals j left-bracket left-parenthesis m double-prime 1 plus m double-prime 2 right-parenthesis omega minus StartFraction m double-prime 1 m double-prime 2 omega cubed Over bold-italic rho bold-italic c squared EndFraction cosine squared theta plus omega bold-italic rho h right-bracket EndLayout  (9.125)

This matrix coefficient is zero at

StartLayout 1st Row 1st Column omega Subscript upper D upper W 2nd Column equals StartRoot StartFraction bold-italic rho bold-italic c squared Over h EndFraction StartFraction m double-prime 1 plus m double-prime 2 plus bold-italic rho h Over m double-prime 1 m double-prime EndFraction EndRoot StartFraction 1 Over cosine theta EndFraction 2nd Row 1st Column Blank 2nd Column almost-equals Underscript rho h much-less-than m double-prime 1 plus m double-prime 2 Endscripts StartRoot StartFraction bold-italic rho bold-italic c squared Over h EndFraction StartFraction m double-prime 1 plus m double-prime 2 Over m double-prime 1 m double-prime 2 EndFraction EndRoot StartFraction 1 Over cosine theta EndFraction EndLayout  (9.126)

The result is the natural frequency of a spring with masses at both ends. The approximation ρhm1+m2 corresponds to a massless spring. However, in double wall construction, the filling of a double wall is lightweight. Figure 9.44 shows the diffuse field transmission loss determined from angle integration. The cavity resonances are smeared out due to the integration but the double wall resonance is still visible.

Figure 9.44 Diffuse field transmission coefficient. Double wall of two limp layers of m1=m2=1 kg/m2 and air cavity of h=10 cm thickness. Source: Alexander Peiffer.

9.5.2 Limp Double Walls with Fiber

As the excellent performance of a double wall is partly compensated by the cavity or double wall resonances, those must be damped by absorbing material, and due to equation (9.122), the characteristic impedance should not be too high. The fiber material of section 9.3.4 is an appropriate candidate for such an application. The air of the double wall cavity is replaced by the fiber material by using the material data of the equivalent fluid. First, we see in Figure 9.45 that the double wall resonance is lower than in air. This results from the lower speed of sound in the fiber material. Second, the transmission loss is much higher than for the air filled cavity, and the notches due to the thickness resonances are not existing. This is a result of the high damping in the fiber material.

Figure 9.45Transmission coefficient for plane waves transmission. Double wall of two limp layers of m1=m2=1 kg/m2 and fiber cavity of h=10 cm thickness. Source: Alexander Peiffer.

It must be kept in mind that this is infinite layer theory. Thus, the calculated transmission loss is slightly overestimated compared to real systems. Every wall must be connected to the other wall by mounts that jeopardize the high isolation; secondly, the finite size of panel and absorber is not taken into account as it is already shown for single walls in section 8.2.4.5. However, for a principle understanding of double wall phenomena, the infinite layer theory is excellent. In Chapters 10 and 11, we will see how the finite size can be considered.

9.5.3 Two Plates with Fiber

Large areas cannot be realized by limp heavy layers, because a certain stiffness is necessary to mechanically hold the plate. The consideration of bending stiffness is done by using the transfer impedance of plates (9.101) in equation (9.119). Figure 9.46 shows a similar shape as for the limp double wall with fiber material. In addition there are the coincidence peaks from the bending waves of the aluminium plates. It is clear that plate materials with similar coincidence frequencies should be avoided, so that only one plate coincides with the exciting or radiating fluid wave.

Figure 9.46Transmission coefficient for plane wave transmission. Double wall of two aluminium plates of thicknesses 3 and 5 mm and fiber cavity of h=5 cm thickness. Source: Alexander Peiffer.

9.5.4 Conclusion on Double Walls

The results from Section 9.5 are some of the most relevant results for passive noise control and acoustic isolation of enclosures or walls. Therefore, we recapitulate the most important conclusions.

Below the double wall resonance, the system behaves as a single wall of the same mass with no benefit regarding noise isolation. Above the double wall resonance, the system performs much better, likely 20 dB and more, because the single transmission coefficients are multiplied and not added.

Thus, design rule number one is: Try to keep the double wall resonance as low as possible. For further details we introduce the reduced mass by

m double-prime Subscript red Baseline equals StartFraction m double-prime 1 plus m double-prime 2 Over m double-prime 1 m double-prime 2 EndFraction  (9.127)

and the frequency is then given by

omega Subscript upper D upper W Baseline equals StartRoot StartFraction bold-italic rho bold-italic c squared Over h m double-prime Subscript red EndFraction EndRoot StartFraction 1 Over cosine theta EndFraction  (9.128)

Weight is a critical parameter in most technical systems. So, we need the best mass distribution for given total mass per area mtot=m1+m2. Let us assume that μ is the fraction of the total mass giving the first mass.

StartLayout 1st Row 1st Column m double-prime 1 2nd Column equals mu m Subscript tot Superscript double-prime Baseline 3rd Column m double-prime 1 4th Column equals left-parenthesis 1 minus mu right-parenthesis m double-prime Subscript tot EndLayout

and the reduced mass reads for μ

m double-prime Subscript red Baseline equals m double-prime Subscript tot Baseline left-parenthesis one-fourth minus left-parenthesis mu minus 0.5 right-parenthesis squared right-parenthesis  (9.129)

The maximum is at μ=0.5 so the lowest double wall frequency is given for a symmetric mass distribution. This ideal condition cannot always be achieved, because the main structure has to provide certain stability and is the heavy part of the wall, for example fuselage panels (Peiffer et al., 2013), walls of buildings, or the body-in-white of a car. In those cases the weight of the second wall is lower than for the main structure m2m1. In this case the reduced mass is approximated by

m double-prime Subscript red Baseline almost-equals m double-prime 2  (9.130)

Figure 9.47 General shape of the double wall transmission and the relationship to main parameters. Source: Alexander Peiffer.

Thus, the light part determines the frequency. As a further consequence it does not make much sense to use a much heavier structure as the first structure, because the resonance is no longer efficiently decreased. So, rule number one under these constraints means to distribute the mass symmetrically if possible. If one wall structure is given, don’t use a mass that is much higher than the first leaf, because this will not be efficient.

For high frequencies the performance according to equation (9.122) is

tau Superscript negative 1 Baseline tilde m Subscript t o t Superscript double-prime 4 Baseline mu squared left-parenthesis 1 minus mu right-parenthesis squared equals m Subscript t o t Superscript double-prime 4 Baseline left-parenthesis left-parenthesis mu minus one-half right-parenthesis squared minus one-fourth right-parenthesis squared  (9.131)

that becomes also maximal at μ=0.5.

The next important parameters are the cavity thickness and stiffness, given by ρc2. Thus, a possible trade off can be space to weight. However, this can be difficult for lower frequencies, because the thickness can, for example, reach more than 15 cm in case of the first blade, passing the frequency of turbo-props around 100 Hz for fuselage noise control.

The stiffness of the centre cavity or layer can be reduced by using fiber material with low speed of sound, but this is partly compensated by the effective higher density. Double walls are also often realized by soft foams with Young’s modulus around E=90200 kN/m2. Those systems are called mass-spring treatments. We should keep in mind that the lowest reasonable Young’s modulus is near 90 kN/m2, because this is the stiffness per area of the air spring in the material, as ρ0c02100 kN/m2.

Rule number two: The space between the layers must be damped. At least there must be some damping; more damping will further increase the performance, but not much. The necessity for damping depends on how much transmission is allowed near the double wall resonance.

Rule number three: Don’t forget the coincidence. If possible use wall properties that lead to different frequencies, or use a lining that is limp.2

To conclude, an efficient double wall system is designed by two panels of similar weight and thick space filled with soft and damped material. One remark concerns the impermeability of the walls. This must be guaranteed to exploit the advantages of double walls. Due the high performance of the total system, small leakages may have a tremendous effect. So, if leakage cannot be avoided due to technical reasons, eg. cable, ventilation, or pipe cut-outs, put special care on the design of sealing systems at those critical areas.

Bibliography

  1. Jean-F. Allard and Noureddine Atalla. Propagation of Sound in Porous Media. Wiley, second edition, 2009. ISBN 978-0-470-74661-5.
  2. Jean F. Allard, Michel Henry, Laurens Boeckx, Philippe Leclaire, and Walter Lauriks. Acoustical measurement of the shear modulus for thin porous layers. The Journal of the Acoustical Society of America, 117(4): 1737–1743, April 2005. ISSN 0001-4966.
  3. M. A. Biot. Generalized Theory of Acoustic Propagation in Porous Dissipative Media. The Journal of the Acoustical Society of America, 34(9A): 1254–1264, September 1962. ISSN 0001-4966.
  4. Yvan Champoux and Michael R. Stinson. On acoustical models for sound propagation in rigid frame porous materials and the influence of shape factors. The Journal of the Acoustical Society of America, 92(2): 1120–1131, August 1992. ISSN 0001-4966.
  5. Frank Fahy. Sound and Structural Vibration: Radiation, Transmission and Response. Academic Press, London, The United Kingdom, 1985. ISBN 0-12-247670-0.
  6. FreeFieldTechnologies. Actran 16.0 User’s Guide. Mont-Saint-Guibert, Belgium, two thousand, sixteenth edition, October 2015.
  7. H. V. Fuchs and X. Zha. Einsatz mikro-perforierter Platten als Schallabsorber mit inhärenter Dämpfung. Acta Acustica united with Acustica, 81(2): 107–116, March 1995.
  8. Kirill V. Horoshenkov, Alistair Hurrell, and Jean-Philippe Groby. A three-parameter analytical model for the acoustical properties of porous media. The Journal of the Acoustical Society of America, 145(4): 2512–2517, April 2019. ISSN 0001-4966.
  9. Harvey H. Hubbard and Acoustical Society of America, editors. Aeroacoustics of Flight Vehicles: Theory and Practice. Published for the Acoustical Society of America through the American Institute of Physics, Woodbury, NY, 1995. ISBN 978-1-56396-407-7 978-1-56396-404-6 978-1-56396-406-0.
  10. Finn Jacobsen. PROPAGATION OF SOUND WAVES IN DUCTS. Technical Note 31260, Technical University of Denmark, Lynby, Denmark, September 2011.
  11. Dah-You Maa. Potential of microperforated panel absorber. The Journal of the Acoustical Society of America, 104(5): 2861–2866, November 1998. ISSN 0001-4966.
  12. Fridolin P. Mechel, editor. Formulas of Acoustics. Springer, Berlin; New York, 2002. ISBN 978-3-540-42548-9.
  13. Clemens Moeser, Alexander Peiffer, Stephan Brühl, and Stephan Tewes. FEM Schalldurchgangsrechnungen einer Doppelwandstruktur. In Fortschritte der Akustik, pages 593–594, Dresden, Germany, March 2008.
  14. M. L. Munjal. Acoustics of Ducts and Mufflers with Application to Exhaust and Ventilation System Design. Wiley, New York, 1987. ISBN 978-0-471-84738-0.
  15. Raymond Panneton. Comments on the limp frame equivalent fluid model for porous media. The Journal of the Acoustical Society of America, 122(6): EL217–EL222, 2007.
  16. Alexander Peiffer, Stephan Tewes, and Stephan Brühl. SEA Modellierung von Doppelwandstrukturen. In Fortschritte der Akustik, Stuttgart, Germany, March 2007.
  17. Alexander Peiffer, Clemens Moeser, and Arno Röder. Transmission loss modelling of double wall structures using hybrid simulation. In Fortschritte Der Akustik, pages 1161–1162, Merano, March 2013.
  18. Allan D. Pierce. Acoustics - An Introduction to Its Physical Principles and Applications. Acoustical Society of America (ASA), Woodbury, New York 11797, U.S.A., one thousand, nine hundred eighty-ninth edition, 1991. ISBN 0-88318-612-8.
  19. Karin Tageman. Modelling of sound transmission through multilayered elements using the transfer matrix method. Master’s thesis, Chalmers University of Technology, Gothenburg, Sweden, 2013.
  20. Huiyu Xue. A combined finite element–stiffness equation transfer method for steady state vibration response analysis of structures. Journal of Sound and Vibration, 265(4): 783–793, August 2003. ISSN 0022-460X.

Notes

  1. 1   The third term corresponds to the moving mass in front of a piston as discussed in section 2.7.3.
  2. 2   This is why the lining is often called trim, because in historic aircraft the second wall was realized by a trimmed membrane
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.139.82.252