Skip to main content
Log in

Analysis and design of acoustic transition sections for impedance matching and mode conversion

  • RESEARCH PAPER
  • Published:
Structural and Multidisciplinary Optimization Aims and scope Submit manuscript

Abstract

This work considers the problem of designing passive transition sections to provide impedance matching and mode conversion for acoustic wave propagation. The base configuration consists of two waveguides connected by a transition section. The objective is to find a placement of material inside this section to make it function as an impedance matcher or a mode converter with minimal losses. A finite element approximation of the Helmholtz equation in a truncated domain together with Dirichlet-to-Neumann type non-reflecting boundary conditions models the wave propagation. Material distribution techniques solve the resulting topology optimization problem and the resulting interfacial devices show good transmission properties.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11
Fig. 12
Fig. 13

Similar content being viewed by others

References

  • Allaire G, Kohn RV (1993) Topology optimization and optimal shape design using homogenization. In: Bendsøe MP, Soares CAM (eds) Topology design of structures. Kluwer Academic Publisher, pp 207–218

  • Ambrosio L, Buttazzo G (1993) An optimal design problem with perimeter penalization. Calc Var 1(1):55–69. doi:10.1007/BF02163264

    Article  MATH  MathSciNet  Google Scholar 

  • Andkjær J, Sigmund O (2011) Topology optimized low-contrast all-dielectric optical cloak. Appl Phys Lett 98(2):3. doi:10.1063/1.3540687

    Article  Google Scholar 

  • Bendsoe MP, Sigmund O (2003) Topology optimization. Theory, methods, and applications. Springer

  • Borel P, Harpøth A, Frandsen L, Kristensen M, Shi P, Jensen JS, Sigmund O (2004) Topology optimization and fabrication of photonic crystal structures. Opt Expr 12(9):1996–2001

    Article  Google Scholar 

  • Borrvall T (2001) Topology optimization of elastic continua using restriction. Arch Comput Methods Eng 8(4):351–385

    Article  MATH  MathSciNet  Google Scholar 

  • Bruns TE, Tortorelli DA (2001) Topology optimization of non-linear elastic structures and compliant mechanisms. Comput Methods Appl Mech Eng 190:3443–3459. doi:10.1016/S0045-7825(00)00278-4

    Article  MATH  Google Scholar 

  • Dobson DC, Cox SJ (1999) Maximizing band gaps in two-dimensional photonic crystals. SIAM J Appl Math 59(6):2108–2120. doi:10.1137/S0036139998338455

    Article  MATH  MathSciNet  Google Scholar 

  • Dobson DC, Wadbro E (2011) Optimization of transmission spectra through periodic aperture arrays. Opt Eng 12(4):509–534. doi:10.1007/s11081-010-9112-4

    Article  MATH  MathSciNet  Google Scholar 

  • Dühring MB, Jensen JS, Sigmund O (2008) Acoustic design by topology optimization. J Sound Vib 317(3–5):557–575. doi:10.1016/j.jsv.2008.03.042

    Article  Google Scholar 

  • Erentok A, Sigmund O (2011) Topology optimization of sub-wavelength antennas. IEEE Trans Antennas Propag 59(1):58–69. doi:10.1109/TAP.2010.2090451

    Article  Google Scholar 

  • Evgrafov A, Rupp CJ, Dunn ML, Maute K (2008) Optimal synthesis of tunable elastic wave-guides. Comput Methods Appl Mech Eng 198(2):292–301. doi:10.1016/j.cma.2008.08.001

    Article  MATH  Google Scholar 

  • Haber RB, Jog CS, Bendsøe MP (1996) A new approach to variable-topology shape design using a constraint on perimeter. Struct Opt 11(1–2):1–12. doi:10.1007/BF01279647

    Article  Google Scholar 

  • Ihlenburg F (1998) Finite element analysis of acoustic scattering. Springer

  • Jensen JS, Sigmund O (2005) Systematic design of acoustic devices by topology optimization. In: Twelfth international congress on sound and vibration, ICVS12 2005, Lisbon

  • Karplus H (1975) Transition section for acoustic waveguides. United States Patent Appl No.: 526,058

  • Kirby R (2008) Modeling sound propagation in acoustic waveguides using a hybrid numerical method. J Acoust Soc Am 124(4):1930–1940. doi:10.1121/1.2967837

    Article  MathSciNet  Google Scholar 

  • Lee JW, Jang GW (2012) Topology design of reactive mufflers for enhancing their acoustic attenuation performance and flow characteristics simultaneously. Int J Numer Methods Eng 91(1):552–570. doi:10.1002/nme.4329

    Article  Google Scholar 

  • Lee JW, Kim YY (2009) Topology optimization of muffler internal partitions for improving acoustical attenuation performance. Int J Numer Methods Eng 80(4):455–477. doi:10.1002/nme.2645

    Article  MATH  Google Scholar 

  • Nomura T, Sato K, Taguchi K, Kashiwa T, Nishiwaki S (2007) Structural topology optimization for the design of broadband dielectric resonator antennas using the finite difference time domain technique. Int J Numer Methods Eng 71:1261–1296. doi:10.1016/j.jcp.2010.05.030

    Article  MATH  Google Scholar 

  • Nomura T, Sato K, Taguchi K, Kashiwa T, Nishiwaki S (2011) A level set-based topology optimization method targeting metallic waveguide design problems. Int J Numer Methods Eng 87(9):844–868. doi:10.1002/nme.3135

    Article  MATH  Google Scholar 

  • Petersson J (1999) Some convergence results in perimeter-controlled topology optimization. Comput Methods Appl Mech Eng 171(1–2):123–140. doi:10.1016/S0045-7825(98)00248-5

    Article  MATH  MathSciNet  Google Scholar 

  • Sigmund O (1994) Design of material structures using topology optimization. Ph.D. thesis. Technical University of Denmark

  • Sigmund O, Jensen JS (2003) Design of acoustic devices by topology optimization. In: Cinquini C, Rovati M, Venini P, Nascimbene R (eds) Short papers of the 5th world congress on structural and multidisciplinary optimization WCSMO5. Lido de Jesolo, pp 267–268

  • Sigmund O, Petersson J (1998) Numerical instabilities in topology optimization: a survey on procedures dealing with checkerboards, mesh-dependencies and local minima. Struct Optim 16(1):68–75. doi:10.1007/BF01214002

    Article  Google Scholar 

  • Svanberg K (1987) The method of moving asymptotes—a new method for structural optimization. Int J Numer Methods Eng 24:359–373. doi:10.1002/nme.1620240207

    Article  MATH  MathSciNet  Google Scholar 

  • Tang SK, Lau CK (2002) Sound transmission across a smooth nonuniform section in an infinitely long duct. J Acous Soc Am 112(6):2602–2611. doi:10.1121/1.1512699

    Article  Google Scholar 

  • Wadbro E, Berggren M (2006) Topology optimization of an acoustic horn. Comput Methods Appl Mech Eng 196:420–436. doi:10.1016/j.cma.2006.05.005

    Article  MATH  MathSciNet  Google Scholar 

  • Yamamoto T, Maruyama S, Nishiwaki S, Yoshimura M (2009) Topology design of multi-material soundproof structures including poroelastic media to minimize sound pressure levels. Comput Methods Appl Mech Eng 198(17–20):1439–1455. doi:10.1016/j.cma.2008.12.008

    Article  MATH  MathSciNet  Google Scholar 

  • Yoon GH (2013) Acoustic topology optimization of fibrous material with delany–bazley empirical material formulation. J Sound Vib 332(5):1172–1187. doi:10.1016/j.jsv.2012.10.018

    Article  Google Scholar 

  • Zhang S (2006) A domain embedding method for mixed boundary value problems. Comptes Rendus Mathematique 343(4):287–290

    Article  MATH  MathSciNet  Google Scholar 

  • Zhou S, Li W, Li Q (2010) Level-set based topology optimization for electromagnetic dipole antenna design. J Comput Phys 229(19):6915–6930. doi:10.1016/j.jcp.2010.05.030

    Article  MATH  MathSciNet  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Eddie Wadbro.

Appendices

Appendix A: Dirichlet-to-Neumann map

To derive the non-reflecting boundary conditions at the truncated waveguides, we consider the setup in a semi-discrete manner as illustrated in Fig. 14. That is, we will treat the problem continuously in the lengthwise direction and discretely in the radial direction. In this derivation, we consider a pipe extending infinitely to the left and assume that we have a discrete piecewise bi-quadratic representation p h of the complex amplitude in the domain S marked grey on the right side. (In the numerical experiments, we have a piecewise bi-quadratic approximation p h of the complex amplitude function p in Ω). Without loss of generality and to simplify notation, it is henceforth assumed that Γ consists of points x = (z,r) with z = 0.

Fig. 14
figure 14

The boundary Γ (dashed line) separates the computational region S (grey) to the right and the semi-discretized infinite pipe to the left. The dash-dotted line is a symmetry line

We first consider a discrete version of the one-dimensional eigenproblem in the radial direction,

$$ \frac{\partial}{\partial r}\left( r \frac{\partial f}{\partial r} \right) = \lambda r f, $$
(23)

to find eigenvalues λ m and eigenfunctions f m . Let \(V_{\Gamma }^{h}\) denote the space of all functions being quadratic on each edge of Γ and let M be the dimension of this space. Letting f denote the coefficient vector of f on Γ, the discrete eigenproblem can be written as

$$ \upgamma\mathbf{K} \mathbf{f} = \lambda \upgamma\mathbf{M} \mathbf{f}, $$
(24)

where γ M and γ K are the mass and stiffness matrices corresponding to problem (23). (Note that, γ can be viewed as a discrete trace operator; for the problem we are interested in, γ M L and γ K L are sub-matrices of M L and K L, respectively, corresponding to the unknowns on Γ L. The boundary matrices M L and K L are defined in (10)). Eigenproblem (24) above has eigenvalues λ m and eigenvectors f m , m=0,1,…,M. The eigenvalues are sorted so that λ i >λ j whenever i>j and that the eigenvectors f m are real and scaled so that

$$ \mathbf{f}_{m}^{\mathsf{T}\!} \upgamma\mathbf{M}\mathbf{f}_{n} = \left\{\begin{aligned} 1 & \text{if}~m=n, \\ 0 & \text{whenever}~ m\neq n. \end{aligned}\right. $$
(25)

Let f m (r) denote the function in \(V^{h}_{\Gamma }\), that has nodal values given by f m . On Γ, the complex amplitude function p h can be written as

$$ p_{h}(0,r) = \sum\limits_{m=0}^{M} G_{m}(p_h) f_{m}(r), $$
(26)

where \(G_{m}(p_h)= \mathbf {f}_{m}^{\mathsf {T}\!} \upgamma \mathbf {M} \upgamma \textbf {p}\) in which γ p denotes coefficient vector of p h | Γ . The solution of the semi-discretized wave propagation problem in the semi-infinite waveguide can be found using separation of variables and is formally written as

$$ p_{h}(z,r) = \sum\limits_{m=0}^{M} f_{m}(r) \left( A_{m} \mathrm{e}^{-\mathrm{i} k_{m} z} + B_{m} \mathrm{e}^{\mathrm{i} k_{m} z} \right), $$
(27)

where A m and B m are complex constants determining the amplitudes of the right and left going waves of mode m, respectively, and \(k_{m}^{2} = k^{2} - \lambda _{m}\) is the reduced wave-number for this mode. The smallest eigenvalue for problem (24) is zero and corresponds to a constant vector (the plane wave mode), hence we have that k 0=k. Evaluating the above expression (27) at z=0 and comparing it with expression (26) implies that G m (p h )=A m +B m . Hence, we define the semi-discrete Dirichlet-to-Neumann type operator on Γ by

$$ \text{DtN}^{\mathrm{h}}(p_h) = -\mathrm{i} \sum\limits_{m=0}^{M} f_{m}(r) k_{m} G_{m}(p_h). $$
(28)

By subtracting the above operator acting on the complex amplitude function p h at Γ from the normal derivative of p h , it follows that

$$\begin{array}{rcl} \frac{\partial p_{h}}{\partial n} &-& \text{DtN}^{\mathrm{h}}(p_h) = \mathrm{i} \sum\limits_{m=0}^{M} f_{m}(r) k_{m} \left( A_{m} - B_{m} \right)\notag\\ & +& \mathrm{i} \sum\limits_{m=0}^{M} f_{m}(r) k_{m} G_{m}(p_h) = 2 \mathrm{i} \sum\limits_{m=0}^{M} f_{m}(r) k_{m} A_m. \end{array} $$
(29)

Thus, by using a boundary condition of the type above, it is possible to set the constants A m determining the complex amplitudes of the incoming waves while perfectly absorbing all outgoing waves. If there is no incoming wave, the operator DtN indeed produces the Neumann data given the Dirichlet data on boundary Γ. The above boundary condition is thus numerically exact in the discrete case, that is, it perfectly absorbs all outgoing waves (the M outgoing wave modes) while specifying the amplitudes of the M incoming wave modes over Γ.

Now, consider the integrals over Γ L and Γ R in problem (8) and let \(\mathbf {f}_{m}^{\mathrm {L}}\) be the mth eigenvector of problem (24) with γ M and γ K replaced by γ M L and γ K L respectively; and let \(\mathbf {f}_{m}^{\mathrm {L}}\) be scaled to satisfy expression (25) with γ M replaced by γ M L. Define \(\mathbf {w}_{m}^{\mathrm {L}}\) as the coefficient vector of \(w^{h}_{m}\in V^{h}\) so that \(\upgamma \mathbf {w}_{m}^{\mathrm {L}} = \mathbf {f}_{m}^{\mathrm {L}}\) and all other components of \(\mathbf {w}_{m}^{\mathrm {L}}\) is zero. In matrix notation, the boundary integral over Γ L in the left hand side of problem (8) can be written

$$\begin{array}{rcl} \int_{\Gamma^{\mathrm{L}}} \varphi_{i} \text{DtN}^{\mathrm{h}}(p_h) = \mathbf{e}_{i}^{\mathsf{T}\!} \mathbf{M}^{\mathrm{L}} \mathbf{DtN}(p_h) \notag\\ &&\qquad = \mathbf{e}_{i}^{\mathsf{T}\!} \mathbf{M}^{\mathrm{L}} \mathrm{i} \sum\limits_{m=0}^{M} k_{m} \mathbf{w}_{m} \left( \mathbf{M}^{\mathrm{L}}\mathbf{w}_{m} \right)^{\mathsf{T}\!} \mathbf{p} \notag\\ &&\qquad = \sum\limits_{j=1}^{N} p_{j} \mathbf{e}_{i}^{\mathsf{T}\!} \underbrace{\left( \mathrm{i} \sum\limits_{m=0}^{M} k_{m} \mathbf{M}^{\mathrm{L}} \mathbf{w}_{m} \left( \mathbf{M}^{\mathrm{L}}\mathbf{w}_{m} \right)^{\mathsf{T}\!} \right)}_{\mathbf{B^{\mathrm{L}}}} \mathbf{e}_j. \end{array} $$
(30)

By letting B R denote a similar matrix corresponding to the integral over Γ R, the discretized variational problem (8) can be written in matrix form as

$$ \left( \mathbf{K}(\alpha_h) - k^{2} \mathbf{M}(\alpha_h) - \mathbf{B}^{\mathrm{L}} - \mathbf{B}^{\mathrm{R}} \right) \mathbf{p} = 2\mathrm{i} k \mathbf{M}^{\mathrm{L}} \mathbf{w}_{0}^{\mathrm{L}}. $$
(31)

Appendix B: Conservation

The derivation below shows that the numerical treatment of the problem is mimetic in that it guarantees conservation of the power in the discrete case.

Before turning to the discrete case, let us first look at sound power of a traveling wave in a cylindrical waveguide in the continuous case. The sound power of this wave is given by the integral of intensity (the product of pressure and velocity) with respect to area. Here, we are studying time harmonic wave propagation in cylindrical symmetry and the sound power for the traveling wave is given by

$$ W = \pi \text{Re} \left\{ \int_{\Gamma} r p\overline{u} \mathrm{d}{\Gamma} \right\}, $$
(32)

where the complex valued functions p and u correspond to pressure and velocity, respectively. By using separation of variables and the relation between pressure and velocity, the functions p and u can be expanded as

$$\begin{array}{@{}rcl@{}} p &=& \sum\limits_{m=0}^{\infty} A_{m} f_{m}(r) e^{-\mathrm{i} k_{m} z}\quad\quad\quad\mathrm{ and } \notag\\ u &=& \sum\limits_{m=0}^{\infty} \frac{A_{m} f_{m}(r) k_me^{-\mathrm{i} k_{m} z}}{\rho c k}, \end{array} $$
(33)

where k m are the reduced wave numbers and the eigenfunctions satisfy a continuous version of (25), that is, they are scaled so that

$$ \int_{\Gamma} r f_{m}(r) \overline{{f}_{n}(r)} \mathrm{d}{\Gamma} = \left\{\begin{aligned} 1 & \text{if}~m=n, \\ 0 & \text{whenever}~m\neq n. \end{aligned}\right. $$
(34)

By inserting expansions (33) for p and u into the expression for sound power (32) and by using the orthonormality of the eigenmodes (34), we get

$$\begin{array}{rcl} W &=& \pi \text{Re} \int_{\Gamma} r p\overline{u} \mathrm{d}{\Gamma} = \pi \text{Re} \left\{ \sum_{m=0}^{\infty} \sum_{n=0}^{\infty} A_{m} \overline{A_{n}}\right. \notag\\ &&\qquad\quad\left.\frac{\overline{k_{n}}}{\rho c k}e^{-\mathrm{i} k_{m} z}\overline{e^{-\mathrm{i} k_{n} z}} \int_{\Gamma} r f_{m}(r) \overline{f_{n}(r)} \mathrm{d}{\Gamma} \right\}\notag\\ &=&\frac{\pi}{\rho c} \text{Re} \left\{ \sum_{m=0}^{\infty} |A_m|^{2} \frac{\overline{k_{m}}}{k} \right\} = \frac{\pi}{\rho c}\sum_{m=0}^{M_{p}}\frac{k_{m}}{k} |A_m|^{2}, \end{array} $$
(35)

where

$$ M_{p} = \max_{m\in \mathbb{N}} m \quad \text{subject to}\,\, k_{m} \in \mathbb{R} $$
(36)

corresponds to the highest possible propagating mode (having a real reduced wave number) in the waveguide. Next, by defining the normalized sound power \(W^{\text {norm}} = \frac {\rho c}{\pi } W\) and letting

$$ P_m=\frac{k_{m}} {k} |A_m|^2 $$
(37)

represent the normalized power of the mth mode wave, we get that the normalized sound power for this traveling wave is given

$$ W^{\text{norm}} = \frac{\rho c}{\pi} W = \frac{\rho c}{\pi}\frac{\pi}{\rho c} \sum\limits_{m=0}^{M_{p}}\frac{k_{m}}{k} |A_m|^{2} = \sum\limits_{m=0}^{M_{p}} P_m. $$
(38)

For the problem studied in this article, the wave propagation in the left waveguide consists of the input wave traveling towards the transition section and the reflected waves, traveling to the left, while the wave propagation in the right waveguide consists of the transmitted waves traveling to the right. Since we have no losses, we have that

$$ W_{\text{refl}}^{\text{norm}} + W_{\text{trans}}^{\text{norm}} = W_{\text{in}}^{\text{norm}}, $$
(39)

where \(W_{\text {refl}}^{\text {norm}}\), \(W_{\text {trans}}^{\text {norm}}\), and \(W_{\text {in}}^{\text {norm}}\) are the normalized power of the reflected wave, the transmitted wave, and the input wave respectively. Boundary conditions (4c) and (4d) ensures that input is planar wave in the left waveguide and specifies the amplitude such that \(W_{\text {in}}^{\text {norm}}=1\) and the power balance (39) can be written as

$$ W_{\text{refl}}^{\text{norm}} + W_{\text{trans}}^{\text{norm}} = 1. $$
(40)

It is perhaps interesting to note that for the chosen input wave, \(W_{\text {refl}}^{\text {norm}}\) is the fraction of input power that gets reflected back to the left waveguide, a fraction commonly referred to as the power reflection coefficient. Similarly, \(W_{\text {trans}}^{\text {norm}}\) is the fraction of input power that gets transmitted to the right waveguide, that is, the power transmission coefficient.

Now, let us turn to the discrete case. The numerical treatment is designed so that it guarantees conservation of power, as shown below. First, making the particular choice \(q_h=\overline {p_{h}}\) in expression (8) yields

$$\begin{array}{rcl} \int_{\Omega} \alpha r \nabla \overline{p_{h}} \cdot \nabla p_{h} - k^{2} \int_{\Omega} \alpha r \overline{p_{h}} p_h -\int_{\Gamma^{\mathrm{L}}} r \overline{p_{h}} \text{DtN}^{\mathrm{h}}(p_h) \notag\\ &&\qquad - \int_{\Gamma^{\mathrm{R}}} r \overline{p_{h}} \text{DtN}^{\mathrm{h}}(p_h) = 2 \mathrm{i} k \int_{\Gamma^{\mathrm{L}}} r \overline{p_{h}} f_{0}^{\mathrm{L}}. \end{array} $$
(41)

The integrands in the domain integrals above are real. By taking the imaginary part of the previous expression and rearrangeing the terms, we obtain

$$\begin{array}{rcl} 0 = \text{Im} &&\left\{ \int_{\Gamma^{\mathrm{R}}} r \overline{p_{h}} \text{DtN}^{\mathrm{h}}(p_h)\right. \notag\\ &&\left. + \int_{\Gamma^{\mathrm{L}}} r \overline{p_{h}} \left( 2 \mathrm{i} k f_{0}^{\mathrm{L}} + \text{DtN}^{\mathrm{h}}(p_h) \right)\right\}. \end{array} $$
(42)

Recall from Appendix A that, in a semi-infinite waveguide and under the assumption that we have discretized the radial direction, the wave propagation can be written as in expression (27) and the DtNh operator is defined in expression (28). Let \(A_{m}^{\mathrm {L}}\) and \(A_{m}^{\mathrm {R}}\) denote the complex constants determining the amplitudes of the incoming (that is moving toward the transition section) waves of mode m at Γ L, and Γ R, respectively. Similarly, let \(B_{m}^{\mathrm {L}}\) and \(B_{m}^{\mathrm {R}}\) denote the complex constants determining the amplitudes of the outgoing waves of mode m at Γ L, and Γ R, respectively.

By the boundary conditions on Γ L and Γ R, we know that \(A_{0}^{\mathrm {L}}=1\), \(A_{m}^{\mathrm {L}}=0\) for all m>0, and \(A_{m}^{\mathrm {R}}=0\) for all m. The constants \(B_{m}^{\mathrm {L}}\) and \(B_{m}^{\mathrm {R}}\) are in the discrete case given by

$$\begin{array}{@{}rcl@{}} B_{m}^{\mathrm{L}} &=& \left( \mathbf{w}_{m}^{\mathrm{L}}\right)^{\mathsf{T}\!}\mathbf{M}^{\mathrm{L}} \mathbf{p} - A_{m}^{\mathrm{L}} \quad \mathrm{ and } \notag\\ B_{m}^{\mathrm{R}} &=& \left( \mathbf{w}_{m}^{\mathrm{R}}\right)^{\mathsf{T}\!}\mathbf{M}^{\mathrm{R}} \mathbf{p} - A_{m}^{\mathrm{R}}, \end{array} $$
(43)

respectively. Here, the notation follows the same style as in Appendix A with the addition of the exponents L and R clarifying whether the object belongs to the left or right waveguide, respectively.

Inserting the series expansions for the complex amplitude function on Γ R into the second integral in expression (42) and expanding the Dirichlet-to-Neumann operator yields

$$\begin{array}{rcl} \int_{\Gamma^{\mathrm{R}}} r \overline{p_{h}} \text{DtN}^{\mathrm{h}}(p_h) \notag \\&&\qquad = \int_{\Gamma^{\mathrm{R}}} \left( \sum_{m=0}^{M^{\mathrm{R}}} f_{m}^{\mathrm{R}}(r) \overline{B_{m}^{\mathrm{R}}} \right) r \left( -\mathrm{i} \sum_{n=0}^{M^{\mathrm{R}}} k_{n}^{\mathrm{R}} f_{n}^{\mathrm{R}}(r) B_{n}^{\mathrm{R}} \right) \notag\\ &&\qquad = - \mathrm{i} \sum_{m=0}^{M^{\mathrm{R}}}\sum_{n=0}^{M^{\mathrm{R}}} k_{n}^{\mathrm{R}}\overline{B_{m}^{\mathrm{R}}}B_{n}^{\mathrm{R}} \int_{\Gamma^{\mathrm{R}}} f_{m}^{\mathrm{R}}(r) r f_{n}^{\mathrm{R}}(r), \end{array} $$
(44)

where M R denotes the number of nodes on Γ R, \(f_{m}^{\mathrm {R}}\) is a piecewise quadratic function on Γ R whose nodal values given by \(\mathbf {f}_{m}^{\mathrm {R}}\) (the mth eigenvector of problem (24) with γ M and γ K replaced by γ M R and γ K R respectively), and \(k_{m}^{\mathrm {R}}\) is the reduced wave number for the mth mode in the right waveguide. By expression (25), the integral above vanishes whenever mn and equals 1 for m=n, thus the above expression simplifies to

$$ \int_{\Gamma^{\mathrm{R}}} r \overline{p_{h}} \text{DtN}^{\mathrm{h}}(p_h) = - \mathrm{i} \sum_{m=0}^{M^{\mathrm{R}}} k_{m}^{\mathrm{R}} |B_{m}^{\mathrm{R}}|^2. $$
(45)

Similarly, inserting the expansions for the complex amplitude function on Γ L into the first integral in expression (42) yields

$$\begin{array}{rcl} &&\int_{\Gamma^{\mathrm{L}}} r \overline{p_{h}} \left( 2 \mathrm{i} k f_{0}^{\mathrm{L}} + \text{DtN}^{\mathrm{h}}(p_h) \right) \notag\\ &&\qquad = \mathrm{i} k - \mathrm{i} k B_{0}^{\mathrm{L}} + \mathrm{i} k \overline{B_{0}^{\mathrm{L}}} - \sum_{m=0}^{M^{\mathrm{L}}} \mathrm{i} k_{m}^{\mathrm{L}} |B_{m}^{\mathrm{L}}|^2. \end{array} $$
(46)

Inserting (45) and (46) into (42) yields

$$\begin{array}{@{}rcl@{}} 0 &=&\text{Im} \left\{ \mathrm{i} k + 2 k \text{Im} \left\{ B_{0}^{\mathrm{L}} \right\} - \mathrm{i} \sum\limits_{m=0}^{M^{\mathrm{L}}} k_{m}^{\mathrm{L}} |B_{m}^{\mathrm{L}}|^2 - \mathrm{i} \sum\limits_{m=0}^{M^{\mathrm{R}}} k_{m}^{\mathrm{R}} |B_{m}^{\mathrm{R}}|^{2} \right\} \notag \\ &=& k - \sum\limits_{m=0}^{M_{\mathrm{p}}^{\mathrm{L}}} k_{m}^{\mathrm{L}} |B_{m}^{\mathrm{L}}|^{2} - \sum\limits_{m=0}^{M_{\mathrm{p}}^{\mathrm{R}}} k_{m}^{\mathrm{R}} |B_{m}^{\mathrm{R}}|^{2}, \end{array} $$
(47)

where

$$\begin{array}{@{}rcl@{}} M_{p}^{\mathrm{L}} &=& \max\limits_{m\in \{1,2,\ldots,M^{\mathrm{L}} \}} m \quad \mathrm{subject~to}~k_{m}^{\mathrm{L}} \in \mathbb{R} \quad \text{and} \notag\\ M_{p}^{\mathrm{R}} &=& \max\limits_{m\in \{1,2,\ldots,M^{\mathrm{R}} \}} m \quad \mathrm{subject~to}~k_{m}^{\mathrm{L}} \in \mathbb{R} \end{array} $$
(48)

correspond to the highest possible propagating mode (having a real reduced wave number) in the left and right waveguide, respectively. Rearranging the terms and dividing by k, the previous expression reduces to

$$ \sum\limits_{m=0}^{M_{\mathrm{p}}^{\mathrm{L}}} \frac{k_{m}^{\mathrm{L}}}{k} |B_{m}^{\mathrm{L}}|^{2} + \sum\limits_{m=0}^{M_{\mathrm{p}}^{\mathrm{R}}} \frac{k_{m}^{\mathrm{R}}}{k} |B_{m}^{\mathrm{R}}|^{2} = 1, $$
(49)

that is, the sum of the outgoing modes square-absolute-amplitudes scaled by their (relative) propagation speed is constant, independent of the material distribution in the transition section. Thus the discrete version of the problem perfectly conserves power. Similarly as in the continuous case above, we let

$$ P_{m}^{\mathrm{L}}=k_{m}^{\mathrm{L}}|B_{m}^{\mathrm{L}}|^2/k, \quad m= 0,\ldots,M_{\mathrm{p}}^{\mathrm{L}} $$
(50)

and

$$ P_{m}^{\mathrm{R}}=k_{m}^{\mathrm{R}}|B_{m}^{\mathrm{R}}|^2/k, \quad m= 0,\ldots,M_{\mathrm{p}}^{\mathrm{R}} $$
(51)

represent the normalized sound power of the mth mode outgoing wave in the left and right waveguides, respectively. (The sound power of the mth mode outgoing wave in the left and right waveguides is given by \(\frac {\pi }{\rho c}P_{m}^{\mathrm {L}}\) and \(\frac {\pi }{\rho c}P_{m}^{\mathrm {R}}\), respectively). Inserting expressions (50) and (51) into expression (49) yields

$$ \sum\limits_{m=0}^{M_{\mathrm{p}}^{\mathrm{L}}} P_{m}^{\mathrm{L}} + \sum\limits_{m=0}^{M_{\mathrm{p}}^{\mathrm{R}}} P_{m}^{\mathrm{R}} = 1. $$
(52)

that is, the same power balance (40) as in the continuous case.

Appendix C: Sensitivity analysis

The method of moving asymptotes requires gradients of each component of the objective function. Below follows a short derivation on how to get the gradient of \(B_{n}^{\mathrm {R}}\), that is complex amplitude of the nth outgoing wave mode in the right waveguide. Fix α h and let δ α h be an arbitrary design variation. Linearizing problem (11), taking into account that K and M are linear in α, yields

$$\begin{array}{rcl} \left( \mathbf{K}(\delta \alpha_h) - k^{2} \mathbf{M}(\delta \alpha_h) \right) \mathbf{p} \notag\\&&\qquad+ \left( \mathbf{K}(\alpha_h) - k^{2} \mathbf{M}(\alpha_h) - \mathbf{B}^{\mathrm{L}} - \mathbf{B}^{\mathrm{R}} \right) \delta \mathbf{p} = 0, \end{array} $$
(53)

where δ p is the first variation of p. In matrix form, \(B_{n}^{\mathrm {R}}\) can be computed as \(B_{n}^{\mathrm {R}} = \left ( \mathbf {w}_{n}^{\mathrm {R}}\right )^{\mathsf {T}\!}\mathbf {M}^{\mathrm {R}} \mathbf {p} - A_{n}^{\mathrm {R}}\) and \(A_{n}^{\mathrm {R}}\) is zero by the boundary condition on Γ R. Differentiating this expression for \(B_{n}^{\mathrm {R}}\) gives

$$ \delta B_{n}^{\mathrm{R}} = \left( \mathbf{w}_{n}^{\mathrm{R}}\right)^{\mathsf{T}\!} \mathbf{M}^{\mathrm{R}} \delta \mathbf{p}. $$
(54)

Letting \(\mathbf {z}_{n}^{R}\) be the solution of problem (11) with right-hand side replaced by \(\mathbf {M}^{\mathrm {R}}\mathbf {w}_{n}^{\mathrm {R}}\), multiplying the linearized state problem (53) with \((\mathbf {z}_{n}^R)^{T}\) and using that the state matrix is symmetric, it follows that

$$\begin{array}{rcl} 0 &=& (\mathbf{z}_{n}^R)^{T} \left( \mathbf{K}(\delta \alpha_h) - k^{2} \mathbf{M}(\delta \alpha_h) \right) \mathbf{p} \notag\\ &&\qquad + (\mathbf{z}_{n}^R)^{T} \left( \mathbf{K}(\alpha_h) - k^{2} \mathbf{M}(\alpha_h) - \mathbf{B}^{\mathrm{L}} - \mathbf{B}^{\mathrm{R}} \right) \delta \mathbf{p} \notag\\ &=& (\mathbf{z}_{n}^R)^{T} \left( \mathbf{K}(\delta \alpha_h) - k^{2} \mathbf{M}(\delta \alpha_h) \right) \mathbf{p} + \left( \mathbf{w}_{n}^{\mathrm{R}}\right)^{\mathsf{T}\!} \mathbf{M}^{\mathrm{R}} \delta \mathbf{p}. \end{array} $$
(55)

From the two above expression, we identify the partial derivatives of \(B_{n}^{\mathrm {R}}\) with respect to the pixel value α E in element E as

$$ \frac{\partial B_{n}^{\mathrm{R}}}{\partial \alpha_{E}} = (\mathbf{z}_{n}^R)_{E}^{T} \left( \mathbf{K}_{E} - k^{2} \mathbf{M}_{E} \right) \mathbf{p}_{E}, $$
(56)

where p E and \((\mathbf {z}_{n}^R)_{E}\) are vectors containing the components of p and \(\mathbf {z}_{n}^{R}\) corresponding to nodes in element E while M E and K E are the element mass and stiffness matrices for element E.

Rights and permissions

Reprints and permissions

About this article

Cite this article

Wadbro, E. Analysis and design of acoustic transition sections for impedance matching and mode conversion. Struct Multidisc Optim 50, 395–408 (2014). https://doi.org/10.1007/s00158-014-1058-2

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s00158-014-1058-2

Keywords

Navigation