• No results found

Chap 8, Waveguide Finite Element Method

N/A
N/A
Protected

Academic year: 2022

Share "Chap 8, Waveguide Finite Element Method"

Copied!
46
0
0

Loading.... (view fulltext now)

Full text

(1)

Waveguide Finite Element Method

HAO LIU, SVANTE FINNVEDEN

MWL, KTH ROYAL INSTITUTE OF TECHNOLOGY

1 EXECUTIVE SUMMARY

In a low frequency analysis of vibration in a built-up structure, the eigen frequencies and the corresponding eigenmodes help the understanding of how and why the structure behaves as it does. These eigen solutions are also useful in calculations of forced response by modal super- position procedures. However, as frequency increases there are more and more modes with increasingly complex forms and the relative information content in each of them decreases.

The level and distribution of damping also becomes increasingly important. For such high frequency motion it is therefore beneficial to change view point and describe the motion in terms of damped waves, in particular so if the structure is relatively long in one direction. If, moreover, it has constant geometrical and material properties along this direction, wave so- lutions can be found with a versatile and numerically efficient method—the waveguide finite element method (FEM).

As an alternative deterministic method to vibro-acoustic problems, waveguide FEM im- proves the efficiency especially for structure who has a waveguide like property. In addition, it provides insightful information on wave solutions, which in turn is used to evaluate para- meters, e.g. group velocity, modal density, for statistical energy analysis (SEA) use. Both of these two characteristics make waveguide FEM a preferable method to bridge the gap between low and high frequencies, which satisfies the requirements of MID-FREQUENCY project to develop effective and efficient vibration and acoustic analysis, modeling and design methods.

In this chapter, the concept and formulation of waveguide FEM is firstly introduced in section 3. In this section, two derivative methods are also described, spectral super ele- ment method (SSEM) and waveguide FEM with Rayleigh-Ritz’s procedure. The derivative techniques complements waveguide FEM in finite length structure and further enhances the efficiency. Subsequently, the performance of waveguide FEM and its derivative methods are demonstrated in section 4. The applications are shown in four different aspects: dispersion curves, finite plate response, curved structure and sound transmission through built-up panels excited by a diffuse sound field. These benchmark illustrations show wide versatility and a good performance; the waveguide FEM should be useful in industrial practices.

(2)

2 INTRODUCTION

A waveguide is a wide-ranging term for a device, which constrains or guides the propagation of mechanical waves along itself. Here it is also assumed that a waveguide has constant geometrical and material properties along one direction.

Waveguide FEM yields equations of motion for systems with wave-propagation along a single direction, in which the structure is uniform. It is then possible to separate the solution to the wave equation into three parts: one part depending on the cross-section, one part depending on the coordinate along the waveguide and one part depending on time.

As an example of a waveguide, a generalized beam, in which longitudinal, torsional, shearing and flexural waves can travel, can be considered. The main idea with a waveguide approach is to study waves propagating in the structure.

The most important benefit with waveguide FE is that it decreases the calculation time compared to conventional FEM since only the cross-section has to be discretized and the number of degrees of freedom is much reduced, especially when the length of the structure is large comparing to the dimension of the cross section. Another advantage compared to conventional FEM, is that it is straight forward to identify and analyze different wave types, which allows a more insightful physical understanding of the structure under investigation.

The ability to handle infinite waveguides, is an additional benefit of this method.

Without external loads, the equations of motion will arrive into a group of homogeneous equations, which could be solved as an eigenvalue problem, shown in section 4.1. Solving the eigenvalue problem, we could have the dispersion curves with wave number against fre- quency. The group velocity and modal density then could be evaluated for further use in energy methods, like Statistical Energy Analysis (SEA).

3 CONCEPT AND FORMULATION OF WAVEGUIDE FEM

When the waveguide FEM describes the free harmonic motion of a structure, it results in a set of coupled ordinary differential equations (ODE). The solutions to these equations are wave functions that describe the wave form and the propagation and decay along the structure as functions of frequency. These solutions are useful for diagnostic purposes, e.g., in Ref. [1]

the size and location of blocking masses was decided based on a waveguide finite elements analysis (FEA) of a wind tunnel and in Ref. [2] a similar analysis of a railway car base frame helped devising a high-frequency and a low-frequency SEA model. Also, the effect of damping treatments is directly inferred from waveguide FEA results [3, 4].

The results from a waveguide FEA can be used in, at least, four different ways to synthe- size the forced response of structures.

• For an infinite waveguide, a spatial Fourier transform of the ODEs gives algebraic equations that are easily solved. The response is then given by an inverse Fourier transform, made by a residue calculus in which the poles are defined by the free wave solutions discussed above [5].

• The damped wave solutions can be used as basis functions in a new, spectral, FE for- mulation for a finite length waveguide. Such a spectral FE is formulated in Ref. [6]

where it is shown that it could be put into an assembly and be combined with other spectral elements and with conventional finite elements, thus providing high computa- tional efficiency in some cases.

(3)

• Certain “convenient” boundary conditions are fulfilled by a wave solution together with a companion wave travelling in the opposite direction. For such boundary conditions, the eigenmodes of a finite length structure are directly identified by a two-dimensional waveguide FEA, which has a much lower computational cost than a conventional three- dimensional FEA. Once the eigenmodes are identified, a modal analysis gives the forced response of the structure. In Ref. [5] the “convenient” boundary condition for a straight waveguide was the shear diaphragm condition, in which all motion within the cross section is blocked while all motion along the waveguide is free.

• In a Rayleigh-Ritz’s procedure, it is possible to directly assume a Fourier series that ful- fill the BCs and then use these as test functions. Since eigenvalue problems are avoided, computational cost is sharply reduced. In addition, in this procedure non-proportional and frequency dependent losses of materials is also possible to be included.

3.1 Waveguide FEM formulation 3.1.1 Displacement and strains

This section presents a derivation of a modified Hamilton’s principle for linear viscoelastic motion upon which a general formulation of waveguide finite elements follows.

The first step in a variational formulation is to explicitly identify the generalized coordi- nates that are varied. Here, an FE displacement formulation is sought and finite elements for isotropic solids and deep shells are considered. The solid’s motion is then given by a three- dimensional displacement field, while the shell’s motion is given by the displacements at its mid-surface plus the rotations about two orthogonal axes on the surface. These displacements are the generalized coordinates and are collected in a vector: u, which is a function of time (t) and location (x).

In linear motion, the strain is given by a linear combination of the displacement and its spatial derivatives. The strain–displacement relations are given by

ε= L (u) (8.1)

whereL is a linear differential operator and the entries to the engineering strain vector ε are the strain components that contributes to the strain energy, ordered in a convenient fashion.

3.1.2 Constitutive equations, strain energy and Hamilton’s principle

The constitutive equation relates the strains to the stresses. For a linear elastic material, the engineering stress vector, σ, is given by

σ= Dε (8.2)

where the rigidity matrix D is symmetric and positive semi-definite. Fung and Tong define the strain energy density function,Uden, such that [7]

σj= ∂Uden/∂εj (8.3)

It follows that the strain energy function is a potential, giving the stresses as functions of the strains. For a linear elastic material it is

Uden=1

TDε (8.4)

(4)

where the superscriptTindicates the transpose of a vector or a matrix.

If the external load on the structure is independent of its motion, as is commonly the case, the potential energy of the loading,A, exists and is a linear function of the displacements:

A = − ˆ

fTudΩ− ˆ

Γ

nTTudΓ (8.5)

where Ω is the structure’s domain, f is the body forces acting on it and Γ is the surface, with outward normal n, where the external traction T is prescribed. On the rest of the boundary there are either homogeneous natural boundary conditions or prescribed displacements. The latter, essential, boundary conditions must be imposed directly on the displacements, since they cannot vary on this part of the boundary.

Upon this basis, Fung and Tong demonstrate, for static problems, the principle of the minimum potential energy.

δ (U + A) = 0 (8.6)

whereδ denotes the first variation and U is the total strain energy of the structure, U =

ˆ

UdendΩ (8.7)

Similarly, for dynamic problems Hamilton’s principle is demonstrated δ

ˆ t2

t1

(U − K + A) dt = 0 (8.8)

where

K =1 2

ˆ

ρ ∂u

∂t

T ∂u

∂t



dΩ (8.9)

is the kinetic energy of the structure andρ is then density.

3.1.3 Viscoelastic material

The stress in a viscoelastic material depends on the strain and also on its history. For a general linear viscoelastic material, the stress is given by a convolution integral as follows [7]:

σ(x, t) = D (ε) = ˆ t

−∞

G(x, t − τ )∂ε

∂τ (x, τ ) dτ (8.10)

where D is a linear operator and the entries to the matrix G are linear combinations of the material’s relaxation functions. Examples of material models that can be written on the form of Eq. (8.10) include generalized Maxwell and Kelvin models [7], fractional derivative models [8] and the Augmented Hooke’s law [9, 10].

Hamilton’s principle does not apply for viscoelastic material. Mathematically speaking, the reason for this is that the operator D is not self-adjoint and because of this, the strain en- ergy functional is not a potential and the variational machinery cannot be used. The operator DAthat is adjoint to D is defined by the following relation:

ˆ

uTD(v) dt = ˆ

vTDA(u) dt (8.11)

(5)

If the Fourier transforms of u and v exist, DAis given by DA(u) =

ˆ

t

−G (x, τ − t)∂u

∂τ (x, τ ) dτ (8.12)

Now, we will have a variational principle for viscoelastic material and therefore need a potential. One option is then the following bi-linear form:

den= 1

4 εAHD(ε) + εTDA εA

(8.13) which, obviously, is self-adjoint. εAis the strain in the adjoint system. This adjoint system is excited by the same forces as the system of interest and is in all aspects identical to this system except for that the material does not obey the constitutive relation given by Eq. (8.10) but

σA= DA εA

(8.14) The strain potential density function ˆUdenhas the property that

δ

ˆ Uˆdendt = 1 2 ˆ

σTδεAdt +1 2

ˆ

σATδεdt (8.15)

and may therefore be used in a variational principle. This modified Hamilton’s principle is given by bi-linear forms, similar to the one in Eq. (8.13). Thus, instead of the work made by external forces in Eq. (8.77), we use the load potential

A = −ˆ 1 2

ˆ

fHu+ fTuA dΩ −1 2

ˆ

Γ

nT Tu + TuA dΓ (8.16) and instead of the kinetic energy in Eq. (8.9), we use the kinetic potential

K =ˆ 1 2

ˆ

ρ ∂uA

∂t

T ∂u

∂t



dΩ (8.17)

Also, we define the total strain potential U =ˆ

ˆ

dendΩ (8.18)

The modified Hamilton’s principle states that δ

ˆ  ˆU − ˆK + ˆA

dt = 0 (8.19)

It is true provided that the displacements obey the equation of dynamic equilibrium and the boundary conditions. The full variational machinery can therefore be used. In particular, Eq. (8.19) is a sound basis for approximate methods, such as the finite element method.

If the responses are harmonic on the form of u(t) = Re ˜ueiωt = 1

2 ˜ueiωt+ ˜ueiωt uA(t) = Re ˜uAeiωt = 1

2 ˜uAeiωt+ (˜u)aeiωt

(8.20)

(6)

The superscriptais introduced forA∗, so thatu˜a = ˜uA

, where denotes Hermitian transpose. Then the modified principle for harmonic response is

δL = δ U − ¯¯ K + ¯A = 0 (8.21)

whereL is denoted the Lagrangian, and the overbar denotes time averaging.

3.1.4 Waveguide finite element discretization

In FEM, the displacements are approximated by polynomial shape functions, which are non- zero within one element only. The amplitude of the shape functions are given by the displace- ments at nodes. An appropriate choice of the shape functions ensure that the displacement fields are continuous between elements, so, they have “compact support” and the functionals in Eq. (8.21) can be evaluated for each of the elements in turn.

In a waveguide FEM model the displacement dependence on the cross-sectional coor- dinates is approximated only. And the displacement dependence on the wave propagation direction could be separated then. An example is given below,

˜

u(x, y, z) = N (y, z) ˜v(x)

˜

ua(x, y, z) = N (y, z) ˜va(x) (8.22) the matrix N are two-dimensional FE shape functions and the entries of the vectorsv and˜ v˜a are the “nodal” displacements on the wave propagation direction, which are functions of the coordinatex.

The strain–displacement relations (8.1) are written in a split operator form and the deriva- tives with respect toy and z are evaluated:

˜

ε= L (˜u) = E0(y, z) ˜v(x) + E1(y, z)∂ ˜v(x)

∂x (8.23)

where the matrices E0and E1contain FE shape functions and derivatives of these, defined by the strain–displacement relations and the FE approximations (8.22).

Upon this basis, and in the absence of external forces, the Lagrangian is given by

L =X

e

ˆ

e

ε˜aTD˜ε˜− ρω2aTu˜ dΩe

=X

e

ˆ

e

! 1 X

m=0 1

X

n=0

meaT

∂xm ETmDE˜ nne

∂xn − ρω2aTe NTN˜ve

"

dΩe

(8.24)

where the summation is taken over the elements, Ωeis the domain of element e and veis the subset of nodes connected to the element. And ˜D is the complex rigidity matrix, given

D˜ = ˆ t

−∞

−iωG (t − τ ) eiω(τ −t)dτ = ˆ

−∞

−iωG (τ ) eiωτdτ (8.25) The elements can be put into an assembly and thus model a great variety of structures. The assembled Lagrangian will be of the form of

L =

ˆ 1

X

m=0 1

X

n=0

m˜vaTe

∂xm Amnne

∂xn − ρω2aTe M˜vedx (8.26)

(7)

where M is a mass matrix and the matrices Amn are generalized stiffness matrices. In the absence of external forces the Lagrangian in Eq. (8.26) is stationary for true motion.

The corresponding Euler–Lagrange equations constitute a set of coupled ordinary differential equations

K22

∂x2 + K1

∂ ˜v

∂x + K0v˜− ρω2M˜v= 0 (8.27) where

K2= −A11, K1= A01− AT01, K0= A00 (8.28) Note, there is a typo in the Reference [11]. The AT10 in Eq. (33) of that paper should be AT01, where A01and A10are anti-symmetric. The matrices in Eq. (8.27) have constant entries and therefore the solutions are exponential functions on the form of

˜

v(x) = ˆveiκx (8.29)

wherev denotes the wave amplitude.ˆ

If Eq. (8.29) is substituted into Eq. (8.27) a twin-parameter eigenvalue problem follows.

Either a linear eigenvalue problem for a given wave order, κ, could be solved or a quadratic eigenvalue problem for a given frequency,ω. The solutions to any of these eigenvalue prob- lems are of interest in their own right and can also be used to synthesis forced response.

3.1.5 Wave solutions and wave numbers

For a given frequency, Eq. (8.27) is a set of coupled ordinary differential equations with con- stant coefficients. Its solution are therefore given by exponential functions and a polynomial eigenvalue problem follows. This problem is then transformed to a standard linear eigenvalue problem, see for example reference [12], with solutions of the form

V(x) = ΦE(x)a (8.30)

where a are wave amplitudes and the entries of the diagonal matrix E(x) are given by (E)ii= eκiix−(κp)iilx (8.31) where κ is a diagonal matrix of eigenvalues. To each component κii thei’th column of Φ gives the corresponding eigenvector. The wave expressions are scaled for reasons of numeri- cal stability by factorsep)iilx, wherelxis half the length of the structure to be investigated and

p)ii= κii, Re[κii] ≥ 0; (κp)ii = −κii, else (8.32) The eigenvalues κii, here also referred to as wave numbers, are solved numerically for a given frequency once the equations of motion have been assembled.

3.2 Spectral FEM formulation

The wave solutions found in the previous section are used as basis functions for a variational principle. Similar to a conventional finite element method, the degrees of freedom are the nodal displacements at the nodes, see Figure 8.1. The nodal displacements at both ends,

(8)

Figure 8.1: Geometry of rectangular element with nodes. Shown with dotted lines are the plate strips used to find the basis functions. Shown to the right are the nodal degrees of freedom.

which are unknowns at this stage, are projected into the whole domain by the wave solutions.

Once all the displacements of the structure are expressed, they could be computed by applying variational principles. In this section, formulation of thin plate spectral finite elements are demonstrated.

3.2.1 Element displacement functions

The wave solutions, V(x) = ΦE(x)a, derived in the previous section will be used to de- scribe the displacement. Now, suppose that displacements Wi of the nodes at the ends are prescribed

B1V(−lx) = W1, B2∂V (−lx)

∂x = W2

B3V(+lx) = W3, B4∂V (+lx)

∂x = W4

(8.33)

where Biare matrices consisting of rows, each filled with zeros except for a one to select the degree of freedom to be restricted. From Eq. (8.33) it is possible to solve the wave ampli- tudes a as functions of the prescribed nodal displacements and thus write the displacement functions as

V(x) = ΦE (x) PW (8.34)

where P is found from solving the system

BP= I, with B=

B1ΦE(−lx) B2ΦκE(−lx)

B3ΦE(+lx) B4ΦκE(+lx)

, W=

 W1 W2 W3

W4

(8.35)

I is an identity matrix of appropriate dimensions. Often, many degrees of freedom are used in order to obtain the wave solutions. Then, similar to a modal substructure, only few degrees of freedom are used in the global model. Thus the number of waves is greater than the number of prescribed boundary conditions, i.e. vector a has more components than vector W. This in turn means that the system is under determined and there are infinitely many solutions. If not otherwise stated, this system is solved here in a least squares sense to the under determined system (see command ‘\’ in MATLAB). The effective rank B of B is determined from an

(9)

orthogonal-triangular decomposition with pivoting and a solution for P is computed which has at mostB non-zero components per column.

If it was possible, when deciding among the available wave solutions from which to find the true displacement of the plate, to select those that are the most likely to contribute to this displacement, the predictions would become more accurate. With this knowledge, it is natural to sort the waves according to their transverse mode shapes and then neglect the highest order mode shapes and possibly also the ones that are not excited by the force. These higher order mode shapes are usually not accurate as they will have less than or close to two nodes per wavelength. The sampling theorem tells us that these waves will then look similar at the nodes to lower order waves and sometimes they replaced lower order waves, when projecting the wave solutions onto the nodal degrees of freedom. This in turn resulted in a decrease in accuracy that was noticeable especially at very low frequencies.

In light of the previous discussion three approaches were used in order to find matrix P.

• Straight-forward approach: All waves found were used.

• Weighted least squares approach: All eigenvectors, described by the columns of Φ, were scaled to have al2-norm equal to1/ |κii|2, where κiiis the corresponding wave number to each eigen vector. This way an important weighting function is introduced, before the system of Eq. (8.35) is solved with a least squares method. Wave solutions that have large wave numbers, describing a rapid decay, will become discriminated, when projecting the solutions onto the nodal degrees of freedom. Instead solutions with small wave numbers, corresponding to waves that have cut on or are about to cut on, will be chosen to describe the displacement of the plate, thereby encouraging the solution to follow Saint-Venant’s principle [7].

• Dynamic condensation approach: A reduced set of waves were used in combination with a dynamic condensation procedure. The main idea behind this is that instead of solving an under determined system in a least squares sense to find P, the virtual work at the boundary is included in the Lagrangian. This results in a new system of equations that can be solved by Gaussian elimination alone.

3.2.2 Dynamic stiffness matrix

The element displacement functions are now described by Eq. (8.34) and from symmetry of the bi-linear Lagrangian the complex conjugate of the displacement functions for the adjoint system are similarly given by

Va(x) = ΦE (x) PWa (8.36)

Substituting these displacement functions into the Lagrangian (8.26) yields L =´lx

lx

2

P

m=0 2

P

n=0

WaTPT ∂m∂xETm(x)ΦTAmnΦn∂xE(x)n PW−

ρω2WaTPTET(x)ΦTMΦE(x)PW dx,

(8.37)

Note that the m’th derivative of the diagonal matrix E(x) is another diagonal matrix κmE(x). For two diagonal matrices d1and d2and another arbitrary matrix v of appropriate dimension the following holds

d1v d2= v. ∗ (diag(d1)diag(d2)T) (8.38)

(10)

where the operator diag produces a column vector from its argument’s main diagonal and.∗

denotes element wise multiplication (as in MATLAB). Using identity (8.38), the Lagrangian (8.37) can be rewritten as

L = WaTDW (8.39)

where the dynamic stiffness matrix D is given by

D = PT(Θ. ∗ EI) P (8.40)

where

Θ=

! 2 X

m=0 2

X

n=0

κmTAmnΦ)κn − ω2Tm00Φ)

"

(8.41)

and

EI= ˆ lx

lx

(diag (E (x))) (diag (E (x)))Tdx (8.42)

The integral in equation (8.42) is solved analytically and consequently the entries of the matrix generating function EIare given by

(EI)ij=(eκii(+lx)−(κp)iilx)(eκjj(+lx)−(κp)jjlx)

− (eκii(−lx)−(κp)iilx)(eκjj(−lx)−(κp)jjlx)/ ((κ)ii+ (κ)jj) (8.43) With the scaling introduced in Eq. (8.32) this expression can be evaluated for largeκ while a Taylor expansion is applied for small values ofκii+ κjj.

The dynamic stiffness matrix (8.40) does not depend on the excitation of the structure.

An arbitrary forcing p can be considered by the inclusion of the virtual work of this force in the Lagrangian.

Lf = ˆ lx

lx

ˆ ly

ly

−pTu− uaTpdy dx = −FTW− WaTF (8.44)

where F is a generalized force vector. For point forces at node locations, the integral is trivial to evaluate. However, for distributed excitation, a particular solution to the non-homogeneous equations of motion has to be included in the formulation and the above integral has to be calculated explicitly, see Ref [13]. Requiring the first variation with respect to the nodal displacements of the adjoint system Wa to be zero in Eq. (8.39) and (8.44), a system of equations for the nodal displacement W is found

DW = F (8.45)

Solving Eq. (8.45) gives the nodal displacements W and then the displacement of the structure at any position could be obtained by Eq. 8.34.

(11)

3.3 Waveguide FEM with convenient BCs

The motion of structures that have constant material and geometric properties along one di- rection can be described by the waveguide FEM [12, 11, 14]. Thus, the harmonic motion displacement field can be defined by 2-dimensional (2D) finite element (FE) shape functions on the cross section and nodal amplitudes that are functions of the coordinates in the wave propagation direction. The discretization leads to a set of coupled ordinary differential equa- tions (ODE) that have constant coefficients.

The solutions to the coupled ODEs are exponential functions which describe the wave propagation and decay along the structure. Most boundary conditions (BC) at the ends of the waveguide are such that if a wave impinges on the boundary a near field is generated and different kinds of waves are reflected. Some convenient boundary conditions, however, are such that there is no near field generation and the only reflected waves are mirrored ones of the incoming waves.

These convenient BCs are here called perfect reflection boundary conditions, and when they apply it is possible to define the nodal amplitudes of finite length waveguide by trigono- metric Fourier series. The set of ODEs turns into algebraic equations and 3-dimensional (3D) forced response calculations are readily made by solving a series of equation system based on a 2D model. The resulting matrices are smaller and have narrower bandwidth, compared to the corresponding 3D matrices. The gain in computer memory requirements and compu- tational speed is therefore substantial.

The present work calculates the STL of structures that are described by the waveguide FEM and that fulfills perfect reflection BCs. This is possible for structures modelled by any of the elements in Refs [12, 11, 14] and also for the rather special elements in Ref [15].

The calculation procedure is an analogy to measurement procedures in a STL suit. Thus, the sound field in the source room is diffuse, defined by a set of plane waves; the receiving room is as if infinite and the structure is surrounded by a rigid baffle. The conditions at the interface between the structure and the baffle are not fully known and may therefore just as well be modelled by the perfect reflection BCs at the waveguide ends, while they can be chosen at will at the other two edges.

Moreover, the calculations are based on a one-way approach where the fluid loading on the structure, both on the source side and the receiving side is neglected. Any fluid struc- ture interaction inside the structure is however modelled as fully coupled. For example, the acoustic motion in the cavity of a double wall partition is fully coupled to the wall vibrations.

3.4 Waveguide finite element formulation with Rayleigh-Ritz procedure 3.4.1 A general formulation of waveguide finite elements

In Ref [11], Finnveden and Fraggstedt describe a derivation of a modified Hamilton’s princi- ple for linear viscoelastic motion by introducing an adjoint system to make a dissipative sys- tem self adjoint as described by Morse and Feshbach [16, P.298-301]. It also shows the for- mulation of general waveguide finite elements. The Lagrangian of a waveguide, aligned with they-axis, with displacements discretized by the assumption of ˜u(x, y, z) = N (x, z) ˜v(y) is given by Eq. 8.26.

3.4.2 Fourier series expansion

The Euler-Lagrange equations that correspond to the homogeneous and undamped system has eigen solutions on the form of

(12)

˜

v= Ψeiκy (8.46)

where the eigen vector Ψ defines the waveforms and the eigenvalue κ defines the propagation along the structure. Ref [16] notes that if Ψeiκyis a solution to the homogeneous equations of motion, so is Ψeiκy, and also

˜

v= Re [Ψ] cos κy − Im [Ψ] sin κy (8.47) Direct inspection of the element formulations and calculated eigen vectors shows that the entries of Ψ can be divided into two groups, see also Gavri´c [17]. The first group contains displacements in the plane of the cross section (thex-z plane) and sound pressures, while the second group contains displacements along the waveguide (they-direction). The entries in one group are all in phase or anti-phase while the two groups are90out of phase.

Upon these arguments follow that nodal displacements, Equation 8.47, not only fulfill the equations of motion but may also fulfill certain BCs. This would happen if the BCs were that the response of one group is zero and the other group is free. These convenient BCs are called perfect reflection conditions, as they may be fulfilled by one incident and a reflected wave.

Specifically, for the shear diaphragm condition the in plane displacement and the sound pressure are zero and the displacement along the waveguide is free; for the gliding BCs the conditions are reversed.

If a waveguide has either of these convenient BCs at both ends, the nodal displacement are given by a Fourier series expansion,

˜

v(y) =X

n

Tnˆvn (8.48)

where the entries of the diagonal matrix Tn are either cos kny or sin kny, depending on whether the corresponding entry of ˆvnbelongs to group one or group two.kn= nπ/l where l is the length of the waveguide.

The displacement, Equation 8.48, is inserted into the Lagrangian, Equation 8.26, and the integral along the waveguide is evaluated. The trigonometric functions are orthogonal and the Lagrangian becomes

L = l 2

X

n

(# P X

p=0 Q

X

q=0

ˆ

vaHnKpnApqKqnˆvn

$

− ω2ˆvaHn KnMKnˆvn+ ˆWn )

(8.49)

where the elements of the diagonal matrix Knarekn, if the corresponded element in Tn is sin kny, and −knotherwise.

3.4.3 Waveguide thick plate and acoustic elements and their couplings

Thick plate elements [18, P.30-33], acoustic elements [14, 15] and their couplings [19, P.638- 639] are formulated following the Rayleigh-Ritz procedure.

Formulation of thick plate element Because of the efficiency this method posses, it will be used to compute the response even at high frequencies. Therefore, the Mindlin plate theory, which includes the shear deformation and rotary inertia effects, is used to formulate the element.

(13)

Displacement field and the ansatz The bending behaviour of the plate is considered, so far, for simplicity. So there will be three DOFs, correspondinglyw displacement along z - axis,θxrotation aroundx - axis, and θy rotation aroundy - axis. And, according to the Mindlin plate theory, the displacement field is given

U(x, y) =

u + zθy(x, y) v − zθx(x, y)

w(x, y)

 (8.50)

where U is the displacement field vector.

We assume the problem is linear, thus the displacements onx - direction, or the cross section, will propagate alongy - axis. The two edges of the plate where y = ylandy = yr are simply supported, which is also called a “convenient” boundary conditions in this report because of the convenience it brings . Thus, the boundary conditions are fulfilled for displacement field on the following form.













u(x, y) = ˜u(x) sin kny v(x, y) = ˜v(x) cos kny w(x, y) = ˜w(x) sin kny θx(x, y) = ˜θx(x) cos kny θy(x, y) = ˜θy(x) sin kny

(8.51)

wherekn = nπ/Lyis the wave number in they-direction, and Lyis the length of the plate in this direction.

Strain and stress The strain vector is given below as

ε=

 εxx

εyy

γxy

γxz

γyz

=

∂u

∂x+ z∂θ∂xy

∂v

∂y − z∂θ∂yx z∂θ∂yy − z∂θ∂xx +∂v∂x+∂u∂y

θy+∂w∂x

−θx+∂w∂y

(8.52)

and the constitutive relation for plane stress problem

D=

E Eν 0 0 0

Eν E 0 0 0

0 0 G 0 0

0 0 0 Gs 0

0 0 0 0 Gs

(8.53)

with

E= E

(1 − ν2) and G = E

2 (1 + ν) and Gs = κG (8.54) whereE is Young’s modulus, ν is Poisson’s ratio, G is the shear modulus, and κ is the modification factor for the non-uniform deformation profile across the thickness of the plate, here taken asκ = 5/6. Hence the strain could be expressed as

σ= Dε (8.55)

(14)

The Lagrangian The Lagrangian is in the form of

ˆ t 0

[δ (T − U ) + δW ] dt = 0 (8.56)

whereT is the kinetic energy, U is the strain energy, and W is the virtual work. They will be derived separately as described below:

U = 1 2 ˆ

εTDεdΩ

= 1 2 ˆ

h(ε00+ zεoz)T+ (ε10+ zε1z)Ti

D[(ε00+ zεoz) + (ε10+ zε1z)] dΩ

= A00+ A01+ A10+ A11

(8.57)

with

























A00= 1 2

ˆ

S

T0000dS +1 2 ˆ

S

h3 12εT

0z0zdS A01= 1

2 ˆ

S

T0010dS +1 2 ˆ

S

h3 12εT

0z1zdS A10= 1

2 ˆ

S

T1000dS +1 2 ˆ

S

h3

12εT1z0zdS A11= 1

2 ˆ

S

T1010dS +1 2 ˆ

S

h3

12εT1z1zdS

(8.58)

where ε00, ε0z, ε10, and ε1zare some components of the strain vector ε































 ε00=

∂u

∂x0

∂v

∂x

θy+∂w∂x

−θx

and ε0z=

∂θy

∂x0

∂θ∂xx 0 0

ε10=

 0

∂v

∂y∂u

∂y

0

∂w

∂y

and ε1z=

 0

∂θ∂yx

∂θy

∂y

0 0

(8.59)

in which, the first sub notation means the order derivative ony, and the second one means whether it timesz or not.

Inserting the ansatz of Equation 8.51 into the strain vector components in Equation 8.59, and then substitute it into Equation 8.58, which turns out that all the terms either have cos2kny or sin2kny. The integration of them on y gives a constant of Ly/2. Therefore the strain energy could be further described as

(15)

























A00= 1 2 ˆ

x

Ly

2 h ˆεooTDεˆoodx + 1 2 ˆ

x

Ly

2 h3

12εˆozTDεˆ0zdx A01= 1

2 ˆ

x

kn

Ly

2 h ˆεooTDεˆ1odx + 1 2 ˆ

x

kn

Ly

2 h3

12εˆozTDεˆ1zdx A10= 1

2 ˆ

x

kn

Ly

2 h ˆε1oTDεˆoodx + 1 2 ˆ

x

kn

Ly

2 h3

12εˆ1zTDεˆ0zdx A11= 1

2 ˆ

x

k2nLy

2 h ˆε1oTDεˆ1odx + 1 2 ˆ

x

k2nLy

2 h3

12εˆ1zTDεˆ1zdx

(8.60)

where

































 ˆ ε00=

∂ ˜u

∂x

0

∂ ˜v

˜ ∂x

θy(x) +∂ ˜w(x)∂x

− ˜θx(x)

and εˆ0z=

∂ ˜θy(x)

∂x

0

∂ ˜θ∂xx(x) 0 0

ˆ ε10=

 0

∂ ˜v

∂ ˜∂yu

∂y

0

˜ w (x)

and εˆ1z=

 0 θ˜x(x) θ˜y(x)

0 0

(8.61)

Hereby, the problem is successfully been reduced by one dimension on the discretization.

In consequence the DOFs needed will be at least6N (Ly/λ) times less, according to the rule of thumb in FEM that6 − 8 nodes per wave length, where N is the nodal DOFs, and λ is the corresponding wave length of the frequency considered. Thus when dealing with long prismatic or relatively high frequency problems, this method is very useful.

Additionally, the integration used in they - direction is exact in contrary of the numerical one used in FEM, which improves the accuracy of the computation. Furthermore, the dis- persion pollution that defined by the difference between exact solution and FEM calculation, which might be worsen as the wave number increases, is also reduced with one dimension less interpolated by FEM meshes.

In a similar approach, the kinetic energy is given by

T =1 2

ˆ

ρ ˙U2dΩ

=1 2

ˆ

x

ω2ρLy

2

 h3 12θ˜y

2(x) +h3 12θ˜x

2(x) + h ˜w2(x)

 dx

(8.62)

And the virtual work is shown

δW = ˆ

S

pzδwdS (8.63)

which is in a simplified form. For different excitations, the work done will be discussed in the sections later.

(16)

3.4.4 Acoustic and thick plate coupling element

Formulations describing a structures coupling to a fluid which is described with a fluid po- tential is presented in e.g. Refs. [20, 21]. The method presented here has been used in Ref.

[14] and has the advantages of describing the undamped coupled system with Hermitian ma- trices and also making the mass matrix, e.g. the matrix proportional to the squared frequency, positive definite.

For a fluid surrounded by a shell, a modified Hamilton’s principle may be written as

δ (Uf− Tf) − δWf = 0 (8.64)

whereδ symbolizes first variation. Ufis the potential andTfis the kinetic energy in the fluid.

δWfsymbolizes the virtual work on the fluid. The corresponding equation for the thick plate is equivalent to Equation 8.64 given

δ (Up− Tp) − δWp= 0 (8.65)

whereUpis the potential andTpis the kinetic energy in the plate.δWpsymbolizes the virtual work on the plate. The virtual work in both plate and fluid is separated due to the coupling between the fluid and the plate. Now, any functional defined by a linear combination of Equation 8.64 and 8.65 defines a functional for the combined system. Here, Equation 8.65 is subtracted from Equation 8.64, written as

δLp+ δLf+ δBc= 0 (8.66)

where

δLp= δ (Up− Tp) (8.67)

is a bilinear functional for the plate

δLf = −δ (Uf− Tf) (8.68)

is a bilinear functional for the fluid and

δBc= δWf− δWs (8.69)

is functional representing the boundary coupling. δLsis used in the derivation of the plate elements.δLfis used in the derivation of the fluid elements andδBcis used for the deriving the fluid-plate coupling elements.

Details of the boundary coupling term,δBc, are explored in the following. To begin, the virtual work on the shell from the fluid is given by

δWp= ˆ

S

δwpdS (8.70)

whereδw, is the virtual displacement of the plate out of the fluid. S is the wetted surface of the fluid plate boundary andp is the fluid pressure. Correspondingly, the virtual work on the fluid from the plate is given by

δWf = − ˆ

S

δpwdS (8.71)

Energy flow into the plate is defined to be positive by Equation 8.70. The energy flow into the fluid is opposite to that into the plate, which explains the minus sign in Equation 8.71.

(17)

3.4.5 Acoustic element

Pressure release boundary condition is assumed at the ends. The reason using this ansatz is to be orthogonal to the normal displacement of the plate element.

p (x, y, z) =X

n

N(x, z) sin (kny) (8.72)

The variational form of the Lagrangian with the substitution of the ansatz leads to

L = Ly

2 X

n

1 2ρ

ˆ



−1

c2NHN+ 1 ω2

 ∂NH

∂x

∂N

∂x +∂NH

∂z

∂N

∂z



+kn2

ω2(−N)H(−N) dΩ (8.73) 3.5 One-way STL calculation model

3.5.1 Diffuse sound field excitation and partition response

Figure 8.2: Scheme (from [22]) of incident plane wave transmission through a baffled parti- tion.

The diffuse sound field [23] is modelled as the superposition of a number of propagating plane waves with equal amplitude from uniformly distributed directions. A plane wave with incidence anglesθ, to the z-axis, and ϕ, to the x-axis, as shown in Figure 8.2, is expressed as pi(x, y, z) = p0e−i(kxx+kyy+kzz) (8.74) wherep0is the amplitude of the wave;kx,kyandkzare defined as

k0= ω c0

, kx= k0sin θ cos ϕ, ky= k0sin θ sin ϕ, kz= k0cos θ (8.75) wherec0is the speed of sound,k0is wavenumber of the fluid.

For a one-way approach, the fluid loading due to the radiation back into the incident space is ignored. Therefore the pressure load on the incident surface (z = 0), given by Sewell [24], and the work done to the structure can be written as

pinc(x, y) = 2p0e−i(kxx+kyy) and W = ˆ Lx

0

ˆ Ly 0

pincw (x, y) dxdy (8.76)

(18)

whereLx, on the cross section, andLy, along the waveguide, are the width and length of the wet surface which the incident wave act on, andw (x, y) is the displacement in z-direction.

After substitution of the ansatzw (x, y) = P ˆwn(x) sin kny for a pinned BCs at ends of a waveguide into the work done in Equation 8.76, the pressure load on the incident surface becomes

W = ˆ

s

p (x, y) w (x, y) ds

= ˆ

s

2p0e−i(kxx+kyy)X

n

[ ˜wn(x) sin (kny)] ds

= 2p0

X

n

x

e−ikxxn(x) dx ˆ Ly

0

e−ikyysin (kny) dy

$

= 2p0

X

n

(#kn 1 − (−1)neikyLy

−ky2+ k2n

x

eikxxn(x) dx )

(8.77)

where Gaussian quadrature is used to evaluate the remaining integral. In this case, the dis- placementw (x) is discretized in the x-direction, and the exact value of eˆ ikxxis used as part of the weighting function.

There is a singular point of Equation 8.77. To calculate the value at the singular point, L’Hˆopital’s rule is used, and the results whenknapproacheskyis given as

#kn 1 − (−1)neikyLy

−k2y+ kn2

$

kykn

= −i (−1)nknLyeikyLy

−2ky



kykn

(8.78)

where in the numerical calculations, the right hand side expression is applied when|kn− ky| < 1E − 6.

3.5.2 Radiated power from the partition

The response of the structure is determined from the waveguide FEM calculation, and the velocityφ normal to the partition is in the form of

φ (x, y) =Xφˆn(x) sin (kny) (8.79) The far field pressure due to the vibration of a baffled structure is given by Rayleigh [25, p.106-112] as

prad= iωρ0

¨ φ (x, y)

r e−ik0rdxdy (8.80)

whereρ0is the density of the fluid,r is the distance from the vibration point to the receiving location.

The partition is assumed to be located around the centre of a hemisphere with large radius,R, compared to the dimension of the partition. A point on the hemisphere is described in spherical coordinates(R, θ, ϕ), and in Cartesian coordinates it is located at

x = R sin θ cos ϕ , y = R sin θ sin ϕ , z = R cos θ (8.81) Therefore, the distance from a point(xp, yp, 0) on the plate to the receiving point is expressed as

(19)

r =h

(x − xp)2+ (y − yp)2+ z2i1/2

= R

#

1 − 2 sin θ cos ϕxp

R −2 sin θ sin ϕyp

R + x2p+ yp2 R2

$1/2

≈ R − xpsin θ cos ϕ − ypsin θ sin ϕ

(8.82)

Since the radius of the hemisphere is much larger than the dimension of the plate, the square of the distance from the source x2p+ yp2 is much smaller than the square of the radiusR. Therefore, the term that is the ratio between them is ignored. according to the binomial formula whenǫ ≪ 1, the expression in the square root can be expanded to get rid of the square root. Substitute the approximation above into the Rayleigh’s integral, it yields

prad(R, θ, ϕ) = iωρ0

¨ φ (x, y)

r eik0rdydx

≈iωρ0

¨ φ (x, y) eik0(R−x sin θ cos ϕ−y sin θ sin ϕ)

R − x sin θ cos ϕ − y sin θ sin ϕ dydx

=iωρ0

2πRe−ik0R

¨

φ (x, y) ei(kxx+kyy)dxdy

(8.83)

The expression above is for the far field sound pressure, which is applicable if k0

2R x2q+ z2q = x2q+ zq20

π ≪ 1 (8.84)

Substitute the velocity of the partition into Equation 8.83, and the radiated sound pressure at the receiving location is expressed as

prad(R, θ, ϕ) = iωρ0

2πReik0RX

n

#kn 1 − (−1)neikyLy

−k2y+ kn2

ˆ Lx

0

φˆn(x) eikxxdx

$

(8.85) At the point wherekyapproacheskn, Equation 8.85 is singular. In this case L’Hˆopital’s rule is used again as in Equation 8.78.

The pressure field radiated from a plane wave excited plate with dimension0.9 × 0.92 × 0.0013 at the far field is shown in Figure 8.3.

3.5.3 Incident power, radiated power and the STL

The power incident on the partition by a plane wave is given as Πi(θ, ϕ) =

ˆ

x

ˆ

y

1

2Re pincpinc ρ0c



cos θdxdy = 1

0cp20LxLycos θ (8.86) The far field radiated waves are plane waves, therefore the radiated power is described as

Πr(θ, ϕ) = ω2ρ0

2c0

¨ sin θ

X

n

#kn 1 − (−1)neikyLy

−ky2+ kn2

ˆ Lx

0

φˆn(x) eikxxdx

$

2

dϕdθ (8.87)

(20)

(a) Pressure field radiated from a 0.9 × 0.92 × 0.0013m plate excited by plane wave with incident angle(θ = 0, ϕ = 0) at 500Hz

(b) Pressure field radiated from a 0.9 × 0.92 × 0.0013m plate excited by plane wave with incident angle(θ = 1, ϕ = 0) at 500Hz

Figure 8.3: Pressure field radiated from a0.9 × 0.92 × 0.0013m plate excited by plane wave with different incident angles at 500Hz

The sound power transmission coefficient,τ (θ, ϕ), is the ratio of the sound power trans- mitted through a partition to the sound power incident on it, written as

τ (θ, ϕ) = Πr(θ, ϕ)

Πi(θ, ϕ) (8.88)

For diffuse sound field excitation, the averaged transmission coefficient is given by Fahy [26, P.293] as

hτ i =

´π/2 0

´

0 τ (θ, ϕ) cos θ sin θdθdϕ

´π/2 0

´

0 cos θ sin θdθdϕ = 1 π

ˆ π/2 0

ˆ 0

τ (θ, ϕ) cos θ sin θdθdϕ (8.89)

wherehi means the average.

The STL is defined as the reciprocal of the sound power transmission coefficient ex- pressed in decibels,

STL= 10 log10

 1 hτ i



(8.90)

4 PERFORMANCE ILLUSTRATIONS

4.1 Dispersion curves and group velocities and modal density 4.1.1 Dispersion curves

This section describes the waves that propagate in a channel beam, shown in Figure 8.4.

The solutions derived from analytical beam theories are compared to the numerical solution provided by the waveguide-FEM and the regions of validity for the different methods are discussed.

Euler beam For an Euler beam, the wave numbers of the propagating waves with transverse motion in the x and y directions, longitudinal motion and rotational motion are given by

(21)

Figure 8.4: Sketch of channel beam. The stars indicate the locations of the mass and shear centers.

kbx= r

ωq

ρA/EIy, kby=

qωpρA/EIx

kl= ω/pE/ρ, kr= ωpρIr/GIt

(8.91)

whereρ is the density, E is Young’s modulus, A is the cross-sectional area, IxandIyare the area moments about the x- and y-axis calculated for the section center,Ir= Ix+ IyandGIt

is the Saint–Venant rotational rigidity. The waveguide-FE is based on thin shell theory and hence for reason of comparison, the transverse and rotational wave numbers in Eq. (8.91) are evaluated with a Young’s modulusE = E/ 1 − ν2; where ν is the Poisson ratio.

Figure 8.5 compares the wave numbers described by Eqs. (8.91) with those calculated by the simple waveguide-FE model. At lower frequencies, the models agree except for a minor difference in the rotational wave. However, already at 5 Hz the rotational wave numbers differ largely and in a cross-over region so the wave numbers for transverse motion in the y direction. Consequently, simple beam theory does not provide an accurate description of all the waves in the beam, except for very low frequencies, and a more elaborate theory is required, as is discussed in Ref. [27].

Simply supported plate In Figure 8.6 the case of a simply supported plate is investigated.

The resulting dispersion curves using nine strip elements are shown. No damping was used here and the known analytical wave numbers for flexural waves, e.g. [28], are shown as comparison. The analytical and predicted dispersion curves agree very well and with more elements the relative error of the predicted wave numbers decreased; a selected study of convergence is shown in Table 1 for the four first order transverse modes.

(22)

Figure 8.5: Wave numbers. Dots, simple waveguide-FE model; solid line, Euler beam theory, Eq. (8.91). In the low-frequency end, these are from top to bottom:kbx,kby,krandkl.

Figure 8.6: Dispersion curves of simply supported plate. Transverse mode shapes are shown next to the curves. Curve with stars, using 9 assembled element strips for the waveguide formulation; solid, analytical solution.

Table 8.1. Relative error of predicted wave number for plate strip in Figure 2 at 800 Hz Transverse mode Number of strip elements

n 3 9 18

n = 1 1.67e-6 1.22e-8 6.80e-10 n = 2 3.62e-4 3.52e-6 2.13e-7 n = 3 1.94e-2 1.19e-4 7.48e-6 n = 4 2.07e-1 2.37e-3 1.60e-4

References

Related documents

The cG(1)cG(1)-method is a finite element method for solving the incompressible Navier-Stokes equations, using a splitting scheme and fixed-point iteration to resolve the nonlinear

The upshot here, compared to [11] is that the pressure in the crack has its own approximation field, allowing accurate approximation of problems where there is a pressure jump

In this paper, we develop a simple finite element method for simulation of embedded layers of high permeability in a matrix of lower permeability using a basic model of Darcy flow

We discuss the regularity properties of the solution and show that if the fracture is sufficiently smooth the problem solution, restricted to the subdomains partitioning the

The finite element solution for the stationary convection-diffusion equation with and without the use of proposed time- relaxation term is observed and they

With this, convergence has been shown for the modified linear poroelastic equations also in the fluid content formulation, discretized spatially with the SBP-SAT technique

The spectra of ―A sailor‖ recorded in the area of light pigment concentration and taken with a blue filter, looks like French Ultramarine, Cerulean Blue, Cobalt Blue and Indigo..

It was shown that the time windows { that is, the eective number of samples used to compute the spectrum at a certain frequency { for common adaptive methods as LMS, RLS and the