Sunteți pe pagina 1din 36

INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN ENGINEERING

Int. J. Numer. Meth. Engng 2008; 73:15711606


Published online 17 July 2007 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/nme.2132

Optimization of an acoustic horn with respect to efficiency


and directivity

Rajitha Udawalpola, and Martin Berggren


Department of Information Technology, Uppsala University, P.O. Box 337, SE-751 05 Uppsala, Sweden

SUMMARY
We consider the problem of designing an acoustic horn in order to efficiently transmit the incoming wave
energy and favorably distribute the energy in the far field. A finite element solution of the Helmholtz
equation, in planar or cylindrical symmetry, models the wave propagation. The transmission efficiency
is monitored by measuring the back reflections into the feeding waveguide, and the far-field directivity
pattern is computed using an integral expression known from scattering theory. The design problem is
formulated as a non-linear least-squares problem, which is solved using a gradient-based algorithm, where
the gradients are provided by solutions of the associated adjoint equations. The results demonstrate that
this approach can generate horns with almost perfect transmission in a wide frequency band. Due to the
improved transmission properties at the lower-frequency region, the optimization with respect to efficiency
also generates improved far-field directivity patterns, that is, patterns that vary less with frequency. It is
possible to obtain even more uniform directivity patterns by explicitly including directivity requirements
in the optimization. However, those improvements in directivity seem to be associated with a substantial
loss of efficiency. Copyright q 2007 John Wiley & Sons, Ltd.

Received 29 August 2006; Accepted 22 May 2007

KEY WORDS: shape optimization; acoustic horn; adjoint equations; Helmholtz equation; far field; finite
element method

1. INTRODUCTION

Horn-like structures appear in devices for both acoustic and electromagnetic waves. Brass
instruments, sirens, and outdoor loudspeakers contain horns that participate in the generation
and distribution of sound waves. Horns are also found in microwave receivers and transmitters,
used for satellite communications, for instance, and small horn-like devices even appear in optical
devices.

Correspondence to: Rajitha Udawalpola, Department of Information Technology, Uppsala University, P.O. Box 337,
SE-751 05 Uppsala, Sweden.

E-mail: Rajitha.Udawalpola@it.uu.se

Copyright q 2007 John Wiley & Sons, Ltd.


1572 R. UDAWALPOLA AND M. BERGGREN

Outgoing wave

Incoming plane wave

2b
2a

d l
x

Figure 1. The type of horn to be optimized. Actual parameter values are specified in Table I.

A horn is an interfacial device, typically mounted at the end of a feeding waveguide. Often
the horn interfaces to open space, as illustrated in Figure 1, but sometimes another component
follows, such as when a microwave horn feeds a dish antenna. A finite number of modes (often
only one) propagate along the waveguide. A horn attached to the end of the waveguide controls
the reflections back to the waveguide as well as how the wave energy is directed away from
the horn. Different applications put different demands on the two aspectsback reflection and
forward directivity controlof the horn function. For a brass instrument, a combined waveguide
horn system is designed so that the waves are strongly reflected back into the instrument at certain
frequencies, corresponding to a few harmonics of the note that is being played. An ideally efficient
loudspeaker horn, in contrast, does not reflect any part of the wave back into the waveguide within
the operational frequency range. It may be advantageous for a microwave antenna to exhibit a very
narrow directivity toward the far field, whereas it is important to obtain a distribution pattern that
does not change much with frequency for a horn loudspeaker.
Todays computer resources, together with the development of sophisticated numerical algo-
rithms, have made accurate mathematical models available for performance analysis of systems
involving horns. Our aim is to use such models not only to analyze the performance of a given
configuration but also to design, by numerical optimization techniques, the shape of a horn to
achieve a set of prescribed properties.
In the literature, there are surprisingly few reports that concern numerical design optimization
of systems involving horns. For electromagnetic waves, there are a few recent examples involving
design of horns used as antenna feeds [14], and Dobson [5] uses optimization techniques to design
a horn-like structure for high-efficiency transfer of optical signals between a dielectric waveguide
and the surroundings. Regarding acoustic waves, Norelands Ph.D. thesis [6] introduces shape
optimization ideas to the design of a trumpet flare in order to enhance the playability of the
instrument. Methods to design the flare of a loudspeaker horn in order to minimize the back

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1573

reflections or, equivalently, to maximize the transmission efficiency have been developed by the
second author of this article and coworkers [7, 8]. The studies showed that the transmission
efficiency of a horn is very sensitive to the precise shape of the horn flare, and that it is possible by
numerical optimization to design very efficient horns within a substantial frequency range. These
studies considered horns in planar symmetry only, that is, horns with an infinite extension in the
direction normal to the plane, and the far-field directivity properties of the horn were not considered,
but only the total transmission efficiency. The present article extends previous methods to consider
also the far-field directivity properties, which are qualities as important as the raw efficiency of
the horn. In contrast with previous studies [7, 8], here the calculations are performed also for
cylindrical symmetry.
In the present study, the horn layout, in terms of the location of the throat and mouth, is
fixed, and the optimization concerns only the horn-flare shape. The numerical computations use
explicit manipulation of the boundary shape together with a body-fitted mesh that adjusts to the
changing boundary shape. This approach yields a very accurate representation of the geometry,
and the numerical computations can be expected to generate accurate approximations of the sound
pressure field. However, dramatic changes in the layout, or topology, are hard to manage with this
approach. There are acoustic design problems, such as the design of folded pipes or horns, acoustic
labyrinths, and acoustical lenses, in which the general configuration is subject to design. These
more general problems would benefit from a different approach than the one considered here. In
one such approach, Jensen and Sigmund [9, 10] as well as Wadbro and Berggren [11, 12] designed
acoustic devices by allowing sound-hard material to be placed arbitrarily within a given region
in space. Such approaches, known as topology optimization, can generate devices of arbitrary
complexity, at the price of less accurate representation of the boundary shape.

2. GOVERNING EQUATION

We assume that the sound propagation is governed by the linear wave equation
2
* P
2
c2 P = 0 (1)
*t
where c is the speed of sound, P is the acoustic pressure, and  is the three-dimensional Laplacian
operator.
Using the ansatz P(x, t) = { p(x)eiwt }, we find time harmonic solutions of the wave equation
for the complex amplitude function p at frequency . Assuming planar or cylindrical symmetries
(Figure 2), and substituting the ansatz into Equation (1), we find that the amplitude function
satisfies the Helmholtz equation
c2 (r p) + 2r p = 0 (2)
where

1 for planar symmetry
r= (3)
y for cylindrical symmetry
Here, y is the distance from the cylindrical symmetry axis and is the gradient operator in the
plane.

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1574 R. UDAWALPOLA AND M. BERGGREN

Figure 2. All calculations are performed in two space dimensions, assuming either planar (left) or cylindrical
symmetry (right). Note the infinite extension in the front view of the planar horn.

out

R
n dsn
sym sym
in n x

Figure 3. The computational domain.

To prepare for the numerical computations, we truncate the exterior region into a bounded
region denoted by  as shown in Figure 3. The amount of computations can be reduced by half
since the horn is symmetric around sym . A right-going wave is specified at in . The boundary
out is an artificial boundary, where we apply the lowest order EnquistMajda absorbing boundary
conditions [13]. The boundary n consists of sound-hard materials. The design boundary dsn ,
which is a part of n , is moved during the optimization process. For more details on the boundary
conditions, see the discussion by Bangtsson et al. [7].
The above boundary conditions together with Equation (2) yield the state equation

c2 (r p) + 2r p = 0 in  (4a)
 
c *p
i + p + c = 0 on out (4b)
R *n
*p
i p + c = 2iA on in (4c)
*n
*p
= 0 on n sym (4d)
*n

where

2 for planar symmetry
= (5)
1 for cylindrical symmetry

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1575

The boundary condition (4c) means that we impose the amplitude A on the right-going wave in
the waveguide, whereas the left-going wave, caused by reflections in the horn, is left unaffected.

3. OBSERVATIONS

We conduct two sets of experiments. In the first, we minimize the reflections in the waveguide in
order to maximize the transmission efficiency. Given a solution to Equation (4), the mean amplitude
 pin on in for planar and cylindrical symmetries is defined as

1
 pin = r p d (6)
 in
where r is given by expression (3) and  is defined by

= r d
in

a for planar symmetry
= (7)
a 2 /2 for cylindrical symmetry

in which a is half the width (diameter) of the waveguide.


The reflection coefficient at in can be defined as
 p in A
I = (8)
A
where p is the acoustic pressure at angular frequency , computed by solving Equation (4), and
A is the amplitude of the incoming wave at in .

Remark 1
If I = 0, then all the wave energy will be transmitted to the surroundings. A brief argument
follows.

The function p satisfies the Helmholtz equation

c2 (r p) + 2r p = 0 (9)

Multiplying (9) by i p and integrating by parts yields


 
0=c 2
(r p) d + 
i p 2
ir p p d
 
   
*p *p
= c2 ir p ir p d + (c2 i p (r p)
+ 2 ir p p)
d
 *n *n 
 
*p
= 2c 
2
ir p d
 *n

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1576 R. UDAWALPOLA AND M. BERGGREN

where the last equality follows from the fact that 2 and c2 are real and thus p satisfies the
Helmholtz equation (9). We have that * p/*n = 0 on n sym so
  
*p *p
 ir p d + ir p d = 0 (10)
in *n out *n
Inside the waveguide, the solution to Equation (9), neglecting evanescent modes, is simply
p = Aeikxn + Beikxn
where A and B are the amplitudes of the right- and left-going wave in the waveguide, respectively,
k = /c is the wave number, and n is the outward-directed unit normal on in . Differentiating in
the n direction yields
*p
= Aikeikxn Bikeikxn
*n
Hence,
*p 2ikxn )]
ip = k[|A|2 |B|2 + 2i(A Be (11)
*n
on in . Moreover, from boundary condition (4b),
 
*p 1 1
i p = i p ik p = k| p|2 i| p|2 (12)
*n R R
on out . Combining Equations (10)(12) yields

(|A| |B| ) =
2 2
r | p 2 | d (13)
out

where expression (7) defines . Neglecting evanescent modes, we have that B =  p in A. Thus,
if the reflection coefficient (8) vanishes, expression (13) reduces to

1
|A| =
2
r | p 2 | d
 out
(End of Remark 1).

In the second set of experiments, we seek to control the variation of far-field directivity pattern
for different frequencies. We thus observe the magnitude of the far-field directivity at a given angle
 and angular frequency . The boundary out cannot be expected to be located in the far field.
We therefore need an extrapolation procedure in order to estimate the true far-field behavior.
First recall the asymptotic behavior of solutions to the Helmholtz equation [14]. Assume that
the only sound source, in the form of a horn, is confined within a bounded domain enclosed by
(d) (d)
out (Figure 4). Let p satisfy the Helmholtz equation (2) in Rd (d = 2, 3) outside of out . Let x
be a unit vector at an angle  with respect to the symmetry axis in Figure 4 and let  be a positive
scalar. The asymptotic behavior of p as  is given by
  
eik  (2) 1
x) =
p( p () + O (14)
 

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1577

x
(d)
out

^x

(d)
Figure 4. Model used for far-field calculations, where out is a closed boundary
surface enclosing all sources.

for planar symmetry and


  
eik  (3) 1
x) =
p( p () + O (15)
 
for cylindrical symmetry. Moreover, classical methods from scattering theory [14] can be used to
(d)
get exact expressions for p in terms of integrals over the boundary (d) out containing p and normal
We use these expressions, substituting the exact solution p with the numerical
derivatives of p.
solution of Equation (4). This substitution will introduce an error due to the artificial boundary
conditions (4b) imposed on out . However, in our numerical studies, we have been careful to
(d)
check, by varying the distance (R ) to out , that there are only small effects generated by the
truncation of the domain within the frequency band we study.
Substituting the solution of Equation (4) into the classical boundary-integral representation of
the far field, and utilizing boundary condition (4b), the far-field pattern at an angle  can be
expressed as
  
1i ik x x 1
p(2)
(, ) = eik x x + ik + p(x) d(x) (16)
4 k (2) out
R 2R
(2)
where out is the union of out and its mirror image with respect to sym and
  
1 ik x x 1
p(3)
(, ) = eik x x + ik + p(x) d(x) (17)
4 (3)
out
R R
(3)
where out is the sphere generated by rotating out around sym . Here, k is the wave number.
Wadbro [15] details the derivations of expressions (14)(17).
The optimization with respect to far-field directivity aims at reducing the variation of far-field
intensities between different frequencies. We thus observe the difference of intensities of two
angular frequencies  and ref at a given angle  as
(d) (d)
I = | p (, )|2 | p (, ref )|2 (18)

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1578 R. UDAWALPOLA AND M. BERGGREN

4. OPTIMIZATION PROBLEM

A proper choice of design variables is crucial for the optimization to succeed. Based on the findings
of Bangtsson et al. [7], we choose to indirectly specify the orthogonal distance  from a straight
line (Figure 5) by solving the equation

 =  in ref
dsn (19a)

=0 at the end points of ref


dsn (19b)

where differentiation is taken in the tangential direction of ref


dsn .
Thus, the design variable  roughly corresponds to the curvature of the design boundary. The
choice of  as our design variable enforces smooth design updates and continuity at the end points
of design boundary dsn . This choice also allows us to enforce convex design boundaries by
constraining  to be of one sign.
We will consider the reflection coefficient given in expression (8) as a function of the design
variable and the frequency and use the notation I (, ). We use the following objective function
of least-squares type for multi-frequency optimization with respect to efficiency:

1 N
J () = 2 d + |I (, i )|2 (20)
2 ref
dsn
2 i=1

where N is the number of frequencies. The first term in objective function (20) is a Tikhonov
regularization term, which will limit excessive design boundary oscillations for >0.
The objective function for optimization with respect to far-field directivity is also of a least-
squares type. Expression (18) is a function of angular frequency , angle of far-field directivity ,
and design variable , for which we use the notation I (, , ). The objective function we use
for multi-angles and multi-frequencies after applying regularization is


1N 
M
J () = 2 d + |I (, ref )|2 + I (, i ,  j )2 (21)
2 ref
d
2 2 i j

where N is the number of frequencies considered in the optimization, M is the number of angles
per frequency, and
>0 is a weighting factor. Note that while attempting, by optimization, to
obtain even directivity, we need to prevent the trivial local minimum that occurs when the intensity
is zero, due to 100 percent reflection for all frequencies. For this, we added the second term in

ref
dsn
n ref

dsn

Figure 5. The function  is the orthogonal distance to the design boundary


(dsn ) from the reference boundary (ref
dsn ).

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1579

objective function (21) to increase the efficiency at the reference frequency, which amounts to a
type of penalty technique to prevent complete reflection.
A minimizer of objective function (21) yields a horn with low reflection at frequency ref ,
while, at the same time, the directivity patterns are matched as much as possible at frequencies
1 , . . . ,  N and angles 1 , . . . ,  M to the one obtained at ref .

5. DISCRETIZATION

We use the finite-element method to discretize Equations (4) and (19). The numerical solution to
Equation (4) is then given by

I
ph (x) = pi i (x) (22)
i=1

where i is a nodal basis function. We use linear Lagrangian elements, that is, each ph is continuous
and piecewise linear on a triangular mesh. The discretized equation (4) can be expressed in matrix
form as
 
c2
c K M+
2 2
M + icM
out in out
p = 2ic AMin e (23)
R
where p = ( p1 , p2 , .. . , p I )T and e = (1, . . . , 1)T . The
 stiffness and mass matrices K and M have
components K i j =  r i j d and Mi j =  r i j d, respectively. Further Mout , Min ,
 
and Min out denote the matrices with components Miout = out r i j d, Miinj = in r i j d,
 j
and Miinj out = in out r i j d, respectively. Expression (3) defines r , and expression (5)
defines .
Numerical approximations of the functions in Equation (19) is given by

I 
I
h (x) = i wi (x), h (x) = i wi (x) (24)
i=1 i=1

where wi is the one-dimensional linear nodal basis function along ref


dsn . The finite element dis-
cretization of Equation (19) can be expressed in matrix form as
Kref a = Mref g (25)


j = ref wi w j d and Mi j = ref wi w j d, respec-
have components K iref
where Kref and Mref ref
dsn dsn
tively.

5.1. Discrete objective functions


This section specifies the terms in objective functions (20) and (21) in the discrete case.
The mean reflection (6) can be expressed in discretized form as
1
 ph, in = eT Min p (26)

where e = (1, . . . , 1)T .

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1580 R. UDAWALPOLA AND M. BERGGREN

x
2
out

j
x

x^

x mj

Figure 6. Geometry model used in the far-field calculation for planar symmetry. Here, 2out is a circle
with radius R , x is a unit vector in the direction of , x j is a unit vector from the center to node j on
j
the upper half of 2out and vector xm is the mirror image of x j .

The far-field patterns (16) and (17) can be expressed in the discretized form as

ph, () = fT Mout p (27)
where Mout is the same matrix as in Equation (23). The vector f has nonzero values only for indices
(d)
that correspond to vertices on boundary out . For planar symmetry, each nonzero component is
given by
 

j (1 i) 1 ik x x j
j
ik x xm ik j ik x x j j ik x xmj
f = ik + (e +e )+ (x x e + x xm e ) (28)
4 k 2R R
j
where x j is the vector from the origin to vertex j, xm is the vector from the origin to the mirror
image of vertex j with respect to boundary sym , and x is the unit vector in the direction of 
(Figure 6).
For cylindrical symmetry, f can be calculated in the following way. Any nonzero component
is given by
 2  
j 1 ik R (a+b cos ) 1
f = e ik(a + b cos ) + ik + d (29)
4 0 R
where a = cos cos , b = sin sin . Here, [0, ] is the angle between the symmetry axis and
vertex j,  is the angle for which we calculate far-field intensities, and is the angle perpendicular
to the plane as shown in Figure 7. We have assumed that x is located in the plane.
Expression (29) can be rewritten as
  2
eika R ikb 2
j i( +kb R cos ) i( +kb R cos )
f = e d + e d
4 2 0 0

   2 
1 ikb R cos
+ ika + ik + e d (30)
R 0

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1581

(3)
out
x

^
x

Figure 7. Geometry model used in the far-field calculation using expression (17) for cylindrical symmetry.
Here, x is a unit vector in the direction of  lying in the plane, x is a vector from the center to an arbitrary
point on the sphere 3out with radius R , is the angle between x and the symmetry axis, and is the
angle between the projection of x on the cross section g (right figure) and the plane.

The Bessel function of the first kind of order m is given by the integral

im 2 i(mv+t cos v)
Jm (t) = e dv (31)
2 0
Using (31) in (30) yields that
 

j eika R 1
f = kb(J1 (kb R ) J1 (kb R )) + 2 ika + ik + J0 (kb R ) (32)
4 R
The discrete form of reflection coefficient (8) is
 ph, in A
I (g, ) = (33)
A
 () in expression (27) as a function of g, ,
Denote by F (g, , ) the far-field intensity ph,
and . Then, the discrete form of the difference in far-field intensities in expression (18) can be
expressed as
I (g, , ) = |F (g, , )|2 |F (g, ref , )|2 (34)
The discrete form of objective function (20), used for optimizing with respect to efficiency, is
1N
J (g) = gT Mref g + |I (g, i )|2 (35)
2 2 i
When optimizing with respect to far-field directivity, we use the following discrete version of
objective function (21):

1N 
M
J (g) = gT Mref g + |I (g, ref )|2 + I (g, i ,  j )2 (36)
2 2 2 i j

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1582 R. UDAWALPOLA AND M. BERGGREN

5.2. Mesh deformations


The shape of the design boundary dsn changes during the optimization process. When the design
boundary is moved, the mesh inside the domain  needs to be altered to preserve the validity
and quality of the mesh. We choose to smoothly deform the mesh instead of complete or partial
remeshing.
In order to avoid too large deformations, we choose an initial design 0dsn , associated with
an orthogonal displacement 0 of the reference design ref
dsn , that should be closer to the optimal
design than dsn . We triangulate the domain associated with 0dsn to obtain the initial mesh from
ref

which the optimization is started.


Define
X0 = (x01 , . . . , x0N )T
where xi0 is the coordinates of mesh vertex i in the initial mesh. The coordinates of the vertices at
the nth design cycle can be represented by a vector
Xn = (xn1 , . . . , xnN )T (37)
where N is the number of vertices in the mesh and n refers to design cycle n. The optimization
algorithm returns the right-hand side g of Equation (25) at the nth design cycle. The corresponding
value of a is calculated by solving Equation (25). Then, we numerically solve equations

u = 0 in  (38a)
u = g(a)nref on 0dsn (38b)

u=0 on \0dsn (38c)

for u using the finite element method. Here, g(a) is a piecewise linear scalar function that attains
the values of (a a0 ) at node points of 0dsn . The values of u at the node points are arranged in
a vector U, which yields a vector of mesh displacements associated with the current g.
Then we update the mesh coordinates using
Xn+1 = X0 + U (39)
Instead of solving Equation (38) at each design cycle, we choose to precompute numerical
solutions for unit changes of each component of the design variable g, and store the solution as
columns of a matrix A. Then, we can express the relationship between changes g of the design
variables and displacements X of the mesh nodes by a linear relation
X = Ag (40)
Using matrix A to update the mesh vertex coordinates is computationally efficient in these
two-dimensional settings. This strategy would not be competitive in three space dimensions due
to excessive storage requirements.

5.3. Sensitivity analysis


We use an optimization algorithmlsqnonlin from Matlabs optimization toolboxthat utilizes
the least-squares structure of the problem. The routine lsqnonlin needs the objective function to be

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1583

written as a simple sum of squares of real numbers. To obtain this form, we introduce the Cholesky
factorization Mref = RT R, where matrix R is upper triangular, and set g = Rg. The variable g will
be the actual design variable manipulated by the optimization routine.
Objective function (35) can then be expressed in the form

1 T 1N
J (g)
= ( g)
( g)
+ |I (g,
i )|2 (41)
2 2 i

where I (g,
i ) = I (R 1 g,
i ), whereas objective function (36) can be written as

1 T 1 1N 
M
J (g)
= ( g) + |
I (g,
( g) )|
2+ I (g,
i ,  j )2 (42)
2 2 2 i j

where I (g,
i ,  j ) = I (R 1 g,
i ,  j ).
Now, the objective functions can be written in the least-squares form required by lsqnonlin:
1 1
J (g)
= fT f , J (g)
= fT f (43)
2 2
where

f =[ g T , [{I (g,
i )} {I (g,
i )}]]T (44)

in which the two last components are repeated for i = 1, . . . , N , and



f = [ g T , {
I (g,
ref )}, {
I (g,
ref )}, [I (g,
i ,  j )]]T (45)

where the last components are repeated for i = 1, . . . , N and j = 1, . . . , M. The vector f has
(2N + L) components and f has (N M + L + 2) components, where L is the number of design
variables.
We need to supply Jacobians of vectors (44) and (45) to the optimization routine. Thus, we
need to determine the derivatives of f and f with respect to g. The Jacobian needed in the
optimization routine can be expressed as

J = [ I, {g I (g,
i )} {g I (g,
i )}]T (46)

where i = 1, . . . , N , for optimization with respect to efficiency and



J = [ I, {g I (g, ref )}, {g I (g,
ref )}, g I (g,
i ,  j )]T (47)

where i = 1, . . . , N and j = 1, . . . , M, for optimization with respect to far-field directivity patterns.


Now we outline how to compute the gradients g I and g I . By expression (40) and the
fact that g = R1 g, we see that a change g in the design variable causes the change

X = AR1 g (48)

in the mesh node coordinates. This change in coordinates changes the components of the matrices
in Equation (23), which causes a change in the objective function I . Denote by I
X
the objective

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1584 R. UDAWALPOLA AND M. BERGGREN

function I regarded as a function of the mesh coordinates. The chain rule and expression (48)
yields

g I = R1 AT x I
X
(49)

The derivation of x I X
is long, and we will just state the result here and refer to Bangtsson
et al. [7] for a derivation of a similar case. Component k of x I X
is
 
*IX
= 2 ph qh (r k ) d c ( ph ( k qh ) + qh ( ph k ))
2
r d
*xk  

c 2
( ph qh )(r k ) d (50)


where ph is given by expression (22) and


I
qh (x) = qi i (x) (51)
i=1

in which the vector q = (q1 , . . . , q I )T solves the adjoint equation


 
c2 1
c K M+
2
Mout icMin
2 out
q = Min e (52)
R 

By expression (34),

I (g, (g,
, ) = |F (g,
, )|2 |F ref , )|2 (53)

where F (g,
, ) = F (R1 g, , ). The gradient of I , which can be expressed similarly
as in expression (49), is

g I = R1 AT x I
X
(54)

where IX
denotes the objective function I expressed as a function of the mesh coordinates. Let
F, denote the function F
X (g,
, ) in expression (53) as a function of the mesh coordinates.
Then expression (53) can be written as

I
X
= (F,
X
 ) + (F, ) (F,ref ) (F,ref )
2 X 2 X 2 X 2
(55)

and x I
X
can be expressed as

x I
X
= 2[{x F,
X
 }{F, } + {x F, }{F, }
X X X

+ {x F,
X
ref }{F,ref } + {x F,ref }{F,ref }]
X X X
(56)

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1585

The derivation of x F,
X
 is done similarly as x I and yields for component k the expression
X

 
*FX,
= 2 ph z h (r k ) d c2 ( ph ( k z h ) + z h ( ph k )) r d
*xk  

c2 ( ph z h )(r k ) d (57)


where ph is given by expression (22) and


I
z h (x) = z i i (x) (58)
i=1

in which the vector z = (z 1 , . . . , z I )T solves the adjoint equation


 
c2
c2 K 2 M + Mout icMin out z = Mout fh (59)
R
where the nonzero components of f are given by expression (28) or (30).

6. EXPERIMENTAL SETUP

The numerical experiments were carried out for both planar and cylindrical symmetries using
horns with dimensions as given in Table I.
We use two major types of plots to describe the properties of the horns: the plot of reflection
spectrum and the plot of far-field directivity patterns in relative dB. Equation (23) is solved to
find the acoustic pressure ph . The reflection coefficient, defined in Equation (33), is calculated
for a frequency band from 10 to 1500 Hz with 10 Hz step length. The absolute values are plotted
 () is calculated
against frequency to obtain the reflection spectrum. The far-field directivity ph,
as in Equation (27) for angles from  = 0 to 2 and angular frequency . The quantity
 ()|

| ph,
20 log10  ()|
| ph, max

 ()|
where | ph, max is the maximum absolute value of directivity pattern in all frequencies and
angles is plotted against  to obtain the far-field directivity patterns. The values resulting from the
above expression can vary in the range from 0 to . The value zero is achieved at the maximum
of directivity patterns. To clearly present the results, we trim the values that are less than 18 dB
below the maximum value.
Table I. The dimensions, as defined in Figures 1 and 3,
used in the numerical experiments.
a (m) b (m) l (m) d (m) R (m)

0.05 0.3 0.5 0.5 1.2

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1586 R. UDAWALPOLA AND M. BERGGREN

Table II. Data for different mesh resolutions, where n is the


number of nodes, m the number of elements, N the number
of design variables, and h max the maximum element size.
Mesh n m N h max (m)

I 857 1519 14 0.1


II 3232 6076 29 0.05
III 12 539 24 304 59 0.025
IV 49 381 97 216 119 0.012
V 195 977 388 864 239 0.006

Table III. Mesh data for different outer boundary radius R , where
n is the number of nodes, m the number of elements, N the number
of design variables, and h max the maximum element size.
R n m N h max (m)

1.2 12 539 24 304 59 0.025


2 24 165 47 392 59 0.025
3 46 931 92 720 59 0.025

0.3

0.2
y

0.1

0.5 0.4 0.3 0.2 0.1 0


x

Figure 8. The initial shape (solid) and the reference shape (dotted).

In order to determine a suitable mesh size, we computed reflection spectra and far-field directivity
patterns for the sequence of uniformly refined meshes specified in Table II. The results only
marginally changed for meshes finer than mesh III, which is why we chose mesh III for the
optimization studies reported below.
To assess the effect of the artificial boundary conditions on out , we performed analogous
studies when varying the radius R in Figure 3 according to Table III. Only minor effects on the

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1587

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400 1600
frequency

Figure 9. Reflection spectrum for the reference shape in planar symmetry


(solid) and in cylindrical symmetry (dashed).

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400 1600
frequency

Figure 10. Reflection spectrum for the initial shape in planar symmetry
(solid) and in cylindrical symmetry (dashed).

reflection spectrum and directivity patterns were observed, which led us to choose R = 1.2 m for
all numerical experiments reported below.
Figure 9 shows the reflection spectra for the reference shape (ref dsn ) of horns in planar and
cylindrical symmetries. It is clear that the reflections are higher in the horn in cylindrical symmetry.
Figure 10 shows the reflection spectra for the initial shape (0dsn ) of the horn depicted in Figure 8
in both planar and cylindrical symmetry. Also, here the magnitude of the reflection coefficient is

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1588 R. UDAWALPOLA AND M. BERGGREN

6
30

12

18

330

Figure 11. Far-field directivity patterns for the reference shape in planar symmetry at frequencies
400 Hz (solid), 600 Hz (dashed), 800 Hz (dashed dotted), and 1000 Hz (dotted).

6
30

12

18

330

Figure 12. Far-field directivity patterns for the reference shape in cylindrical symmetry at frequencies
600 Hz (solid), 700 Hz (dashed), 800 Hz (dashed dotted), and 900 Hz (dotted).

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1589

higher for the horn in cylindrical symmetry. However, the magnitude of the reflection coefficient
in both cylindrical and planar symmetries is smaller for the initial shape than for the reference
shape. This is because we choose the initial shape to be closer to the optimized horn shape to
avoid large mesh deformations.
Figure 11 shows the far-field directivity patterns in planar symmetry for the reference shape.
A beaming effect is clearly visible; that is, with increasing frequency, the intensity will be more
concentrated in the direction of the horn. Figure 12 shows the far-field directivity for the refer-
ence shape in cylindrical symmetry, and reveals an even stronger beaming effect than the planar
symmetry.

7. RESULTS

7.1. Optimization with respect to efficiency


We performed optimization at single frequencies using planar symmetry. Figures 13 and 14 show
the reflection spectra of the horns optimized for frequencies 400, 600, 800, and 1000 Hz. Figures
15 and 16 show the corresponding designs. No geometry constraints or regularization ( = 0 in
objective function (41)) was applied during the optimization.
Optimization of a horn in planar symmetry, for a frequency band from 350 to 1150 Hz at 31
frequencies, was also performed. Figure 17 shows the reflection spectrum of the optimized horn. It
shows almost perfect transmission properties throughout the frequency band. The design obtained
by this optimization is shown in Figure 18. No geometric constraints, but a small amount of
regularization ( = 105 ) was applied during the optimization. The irregular shape may not be
suitable for manufacturing. Comparing Figures 11 and 19, we see that the beaming effect of the
horn has been reduced. This is because the relative improvement of efficiency at low frequencies
is higher than at high frequencies.

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400 1600
frequency

Figure 13. The reflection spectrum for the horns in planar symmetry optimized for 400 Hz (solid), 600 Hz
(dashed), and the spectrum of the reference shape (dotted).

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1590 R. UDAWALPOLA AND M. BERGGREN

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400 1600
frequency

Figure 14. The reflection spectrum for the horns in planar symmetry optimized for 800 Hz (solid), 1000 Hz
(dashed), and the spectrum of the reference shape (dotted).

0.3

0.2
y

0.1

0
0.5 0.4 0.3 0.2 0.1 0
x

Figure 15. Shape of the horns in planar symmetry optimized for 400 Hz (solid), 600 Hz (dashed),
and the reference shape (dotted).

The same experiment was carried out again, but with added convexity constraints for the de-
sign boundary. Figure 20 shows the reflection spectrum of the optimized horn. The corresponding
shape in Figure 21 clearly shows that the design is smoother than the designed obtained with-
out convexity constraints. The transmission properties throughout the frequency band are not as
perfect as in the optimization without convexity constraints, since we reduce the design space by
applying constraints. However, as a side effect of the smoother shape, notice that the efficiency
for higher frequencies is better with the design obtained with convexity constraints. By comparing

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1591

0.3

0.2
y

0.1

0.5 0.4 0.3 0.2 0.1 0


x

Figure 16. Shape of the horns in planar symmetry optimized for 800 Hz (solid), 1000 Hz (dashed),
and the reference shape (dotted).

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400 1600
frequency

Figure 17. Reflection spectrum of the horn in planar symmetry optimized with no geometry
constraints for a frequency band from 350 to 1150 Hz for 31 frequencies (solid) and the
spectrum of the reference shape (dashed).

Figures 19 and 22, we can see that applying constraints does not considerably affect the far-field
directivity patterns.
We also performed single-frequency optimization for the horn in cylindrical symmetry. We
did not apply regularization ( = 0) or geometric constraints for the optimization. The horn
was optimized at frequencies 400, 600, 800, and 1000 Hz, the same frequencies as used for

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1592 R. UDAWALPOLA AND M. BERGGREN

0.3

0.2
y

0.1

0.5 0.4 0.3 0.2 0.1 0


x

Figure 18. Shape of the horn in planar symmetry optimized with no geometry constraints for a
frequency band of 3501150 Hz (solid) and the reference shape (dashed).

6
30

12

18

330

Figure 19. Far-field directivity patterns for frequencies 400 Hz (solid), 600 Hz (dashed), 800 Hz
(dotted-dashed), and 1000 Hz (dotted), of the horn in planar symmetry optimized with no geometry
constraints for the frequency band 3501150 Hz.

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1593

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400 1600
frequency

Figure 20. Reflection spectrum of the horn in planar symmetry optimized with convexity
constraints for a frequency band from 350 to 1150 Hz at 31 frequencies (solid) and spectrum
of the reference shape (dashed).

0.3

0.2
y

0.1

0
0.5 0.4 0.3 0.2 0.1 0
x
Figure 21. Shape of the horn in planar symmetry optimized with convexity constraints for a frequency
band from 350 to 1150 Hz (solid) and reference shape (dashed).

single-frequency optimization in planar symmetry. The resulting reflection spectra are shown
in Figures 23 and 24. It is clear that reflection can be minimized perfectly for a single fre-
quency. However, when comparing with the same case of planar symmetry, we can see that overall
efficiency is less for the horn in cylindrical symmetry. We can see this behavior in many of our

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1594 R. UDAWALPOLA AND M. BERGGREN

6
30

12

18

330

Figure 22. Far-field directivity patterns for frequencies 400 Hz (solid), 600 Hz (dashed), 800 Hz (dot-
ted-dashed), and 1000 Hz (dotted), of the horn in planar symmetry optimized with convexity constraints
for a frequency band of 3501150 Hz.

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400
frequency

Figure 23. Reflection spectrum for the horns in cylindrical symmetry optimized for 400 Hz (solid),
600 Hz (dashed), and the spectrum of the reference shape (dotted).

experiments. The optimal shapes are shown in Figures 25 and 26. When compared with the shapes
for planar symmetry (Figures 15 and 16), the curvature of the optimized shapes in cylindrical
symmetry is larger.

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1595

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400
frequency

Figure 24. Reflection spectrum for the horns in cylindrical symmetry optimized for 800 Hz (solid),
1000 Hz (dashed), and the spectrum of the reference shape (dotted).

0.3

0.2
y

0.1

0.5 0.4 0.3 0.2 0.1 0


x

Figure 25. Shape of the horns in cylindrical symmetry optimized for 400 Hz (solid), 600 Hz (dashed),
and the reference shape (dotted).

The optimization of the horn in cylindrical symmetry for a frequency band from 350 to 1150 Hz
with 31 frequencies was performed as the next experiment. Tikhonov regularization of = 105 was
applied during the optimization. However, no constraint on geometry was applied. The resulting
reflection spectrum is shown in Figure 27. We could get good transmission properties within the
frequency band. The shape of the design is shown in Figure 28. Again, we see that not imposing

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1596 R. UDAWALPOLA AND M. BERGGREN

0.3

0.2
y

0.1

0.5 0.4 0.3 0.2 0.1 0


x

Figure 26. Shape of the horns in cylindrical symmetry optimized for 800 Hz (solid), 1000 Hz (dashed),
and the reference shape (dotted).

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400
frequency

Figure 27. Reflection spectrum of the horn in cylindrical symmetry optimized with no ge-
ometry constraints for a frequency band from 350 to 1150 Hz at 31 frequencies (solid)
and spectrum of the reference shape (dashed).

any geometry constraints may result in an irregular shape that likely would be complicated to
manufacture. We notice in Figure 29 that the beaming effect is reduced compared to reference
shape (Figure 12).
The same experiment was performed again, but with convexity constraints on the geometry.
The resulting spectrum is shown in Figure 30. We can see that the reflections within the optimized

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1597

0.3

0.2
y

0.1

0.5 0.4 0.3 0.2 0.1 0


x

Figure 28. Shape of the horn in cylindrical symmetry optimized with no geometry constraints for a
frequency band from 350 to 1150 Hz (solid) and the reference shape (dashed).

6
30

12

18

330

Figure 29. Far-field directivity patterns for frequencies 600 Hz (solid), 700 Hz (dashed), 800 Hz (dot-
ted-dashed), and 900 Hz (dotted), of the horn in cylindrical symmetry optimized with no geometry
constraints for a frequency band 3501150 Hz.

frequency band have increased slightly (compare with Figure 27). This is reasonable since we
reduced the design space by applying constraints on the geometry. On the other hand, the smoother
shape has improved the transmission efficiency at higher frequencies that were not included in the

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1598 R. UDAWALPOLA AND M. BERGGREN

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400
frequency

Figure 30. Reflection spectrum of the horn in cylindrical symmetry optimized with con-
vexity constraints for a frequency band from 350 to 1150 Hz at 31 frequencies (solid) and
spectrum of the reference shape (dashed).

0.3

0.2
y

0.1

0
0.5 0.4 0.3 0.2 0.1 0
x

Figure 31. Shape of the horn in cylindrical symmetry optimized with convexity constraints for a
frequency band from 350 to 1150 Hz (solid) and the reference shape (dashed).

optimization. Figure 31 shows the resulting design shape. The far-field directivity patterns for the
design is shown in Figure 32. The beaming effect has reduced compared to the reference shape.
However, applying convexity constraints has made a little difference in the far-field directivity
patterns.

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1599

6
30

12

18

330

Figure 32. The far-field directivity patterns for frequencies 600 Hz (solid), 700 Hz (dashed), 800 Hz
(dotted-dashed), and 900 Hz (dotted), of the horn in cylindrical symmetry optimized with no convexity
constraints for the frequency band 3501150 Hz.

7.2. Optimization with respect to directivity


Next, optimization with respect to far-field directivity patterns was performed for the horn in
planar symmetry. In this experiment, we try to match the far-field intensity, the square of the
absolute values of far-field directivity, in the direction of zero angle at frequencies 600, 800, and
1000 Hz, with the value at 400 Hz. At the same time, we minimize the reflection ratio at 400 Hz (the
reference frequency) in order to prevent 100% reflection. This strategy reduces the difference of
far-field intensities between the reference frequency and other frequencies higher than the reference
frequency. We do not impose any constraints on the geometry, but use regularization with = 105
and weighting factor
= 1 in objective function (42).
The resulting far-field directivity patterns are shown in Figure 33. When compared with the far-
field directivity patterns in reference shape (Figure 11), far-field intensity in the direction of zero
degrees are reduced. Also, compared with the far-field directivity patterns for the horn optimized
for efficiency (Figures 19 and 22), we see that we have achieved very even far-field directivity
properties in the direction of zero degrees. However, for angles other than zero, the directivity
properties are not so even. The resulting shape is shown in Figure 34. The reflection spectrum for
the optimized shape is shown in Figure 35. We have achieved optimum transmission properties at
400 Hz, but the reflections at other frequencies are high.
The same experiment was carried out again, but this time requesting even directivity at three
different angles 0, 25, and 30 . Figure 36 shows the directivity patterns at 400, 600, 800, and
1000 Hz. It is clear that far-field intensities are almost perfectly matched. When compared with
optimization with respect to one angle (Figure 33), the directivity patterns are well matched
in all directions. The shape of the horn shown in Figure 37 is not as smooth as the shape

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1600 R. UDAWALPOLA AND M. BERGGREN

6
30

12

18

330

Figure 33. Far-field patterns at frequencies 400 Hz (solid), 600 Hz (dashed), 800 Hz (dotted-dashed), and
1000 Hz (dotted) for the horn in planar symmetry optimized for far-field pattern for in the direction
of zero degrees for the same frequencies.

0.3

0.2
y

0.1

0.5 0.4 0.3 0.2 0.1 0


x

Figure 34. Shape of the horn in planar symmetry optimized for far-field directivity pattern at
400, 600, 800, and 1000 Hz (solid) and the reference shape (dashed).

obtained when optimizing for one angle (Figure 34). From the reflection spectrum shown in
Figure 38, we see that the horn is less efficient than the one optimized with only one angle
(Figure 35).

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1601

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400 1600
frequency

Figure 35. Reflection spectrum for the horn in planar symmetry optimized for directivity at
400, 600, 800, and 1000 Hz (solid) and spectrum for the reference shape (dashed).

6
30

12

18

330

Figure 36. Far-field patterns at frequencies 400 Hz (solid), 600 Hz (dashed), 800 Hz (dotted-
dashed), and 1000 Hz (dotted) for the horn in planar symmetry optimized for far-field
pattern in the directions of 0, 25, and 30 .

Turning now to horns in cylindrical symmetry, optimization to achieve even far-field directivity
patterns in the direction of zero angle was carried out. The reflection is minimized at 600 Hz
and the far-field intensities at frequencies 700, 800, and 900 Hz are matched with the intensity

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1602 R. UDAWALPOLA AND M. BERGGREN

0.3

0.2
y

0.1

0.5 0.4 0.3 0.2 0.1 0


x

Figure 37. Shape of the horn in planar symmetry optimized for far-field directivity pattern at 400,
600, 800, and 1000 Hz (solid) in directions of 0, 25, and 30 and the reference shape (dashed).

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400 1600
frequency

Figure 38. Reflection spectrum for the horn in planar symmetry optimized for di-
rectivity at 400, 600, 800, and 1000 Hz in the directions of 0, 25, and 30 (solid)
and spectrum for reference shape (dashed).

at 600 Hz. We do not impose any constraints on the geometry, but use regularization with = 105
and weighting factor
= 102 in objective function (42). The far-field directivity patterns for the
optimized horn are shown in Figure 39. Compared with previous directivity patterns (Figures 29
and 32), we have obtained a much more even directivity. The design shape after optimization
is given in Figure 40. The reflection spectrum is given in Figure 41. We have achieved perfect

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1603

6
30

12

18

330

Figure 39. The far-field patterns at frequencies 600 Hz (solid), 700 Hz (dashed), 800 Hz
(dotted-dashed), and 900 Hz (dotted) for the horn in cylindrical symmetry optimized for
far-field directivity in the direction of zero degrees.

0.3

0.2
y

0.1

0.5 0.4 0.3 0.2 0.1 0


x

Figure 40. Shape of the horn in cylindrical symmetry, optimized for far-field directivity patterns at
600, 700, 800, and 900 Hz in the direction of zero degrees (solid) and the reference shape (dashed).

transmission properties at 600 Hz. The efficiency at frequencies other than 600 Hz is, however, not
so impressive. The same experiment was performed adding more angles, but without significantly
improving the evenness in the far-field patterns.

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1604 R. UDAWALPOLA AND M. BERGGREN

0.8

0.6
|R|

0.4

0.2

0
0 200 400 600 800 1000 1200 1400
frequency

Figure 41. Reflection spectrum of the horn in cylindrical symmetry, optimized for directivity
patterns for 600, 700, 800, and 900 Hz in the direction of zero degrees (solid) and reflection
spectrum of the reference shape (dashed).

8. DISCUSSION AND CONCLUSIONS

Our results demonstrate that the transmission efficiency and the far-field directivity properties of
acoustic horns are very sensitive to the shape of the horn flare. A rather limited change in the design
can substantially change the acoustic properties in an unintuitive way. Together, these observations
strongly suggest that numerical shape optimization should seriously be considered as a tool for
practical horn design.
In the present study, we use a very detailed geometry description of the shape. In fact, the number
of design variables always equals the number of mesh points located at the design boundary. A naive
adoption of such a fine-grained parameterization easily results in irregular and mesh-dependent
designs. To mitigate these problems, we adopted the strategy motivated and discussed in detail by
Bangtsson et al. [7], a strategy that promotes smooth design updates at each optimization iteration.
We also added, when needed, a small Tichonov regularization term to the objective function
in order to dampen the occurrence of large design oscillation. Nevertheless, for multi-frequency
optimization, the optimized shapes still turned out to be somewhat irregular (Figures 18 and 28).
The addition of a convexity constraint resulted in much smoother shapes (compare Figures 18
and 28 with Figures 21 and 31) at only a small price in terms of efficiency. Another advantage
with the convexity constraint is that the resulting horn offers improved transmission efficiency at
frequencies higher than those included in the optimization due to the smoother shape.
An algorithmic difference from previous studies [7, 8] is that we utilize better the least-squares
structure of the problem and provide separately the gradients of the real and imaginary parts
of the reflection coefficient to the optimization algorithm. This change, compared with previous
practice, significantly lowered the number of iterations needed by the optimization algorithm.
Another advantage of the least-squares formulation is that the adjoint equation becomes completely

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
OPTIMIZATION OF AN ACOUSTIC HORN 1605

independent from the forward equation in the case of optimization with respect to transmission
efficiency. This independence should be valuable for parallel implementations.
The results demonstrate that we can design horns with almost perfect transmission properties
within a significantly wide frequency band. The improvement in transmission efficiency, compared
with the reference shape, is larger for the horn with cylindrical symmetry, even though the reflections
in the optimized cylindrical horn are slightly higher than in the optimized planar horn. The
improvement in transmission efficiency is particularly pronounced at lower frequencies, since the
efficiency for the initial horn was low in this region. The improved efficiency at lower frequencies
also makes the far-field directivity patterns more even at different frequencies in the optimized horn
compared with the reference horn (compare Figures 22 and 11 for the planar horn, and Figures
32 and 12 for the cylindrical horn), even though directivity properties were not considered at all
in those optimizations! The price to pay for the improved efficiency is the occurrence of small
secondary lobes, particularly in cylindrical symmetry.
Directly optimizing with respect to directivity is more delicate. To obtain meaningful results,
we found it necessary to include also a term involving the reflection in the objective function. We
have demonstrated that it is possible to obtain very even far-field directivity patterns with respect
to frequency (Figures 36 and 39). However, the penalty on the efficiency is high. To obtain more
detailed control over the far-field directivity without too much penalty on efficiency, it may be
necessary to include additional acoustical components, such as an acoustic lens. The combined
hornacoustic-lens system should of course be subject to optimization.

ACKNOWLEDGEMENTS
The authors are grateful to Erik Bangtsson and Daniel Noreland for input regarding their earlier work on
the same configuration. Thanks also to Eddie Wadbro for help with the far-field expressions, and to Pablo
Seoane for knowledgeable feedback from the world of professional audio.

REFERENCES
1. Deguchi H, Tsuji M, Shigesawa H. Compact low-cross-polarization horn antennas with serpentine-shaped taper.
IEEE Transactions on Antennas and Propagation 2004; 52(10):25102516.
2. Granet C. Optimisation of a spline-profile corrugated horn for prime-focus operations. Electronic Letters 2004;
40(9):522523.
3. Granet C, James GL, Bolton R, Moorey G. A smooth-walled spline-profile horn as an alternative to the
corrugated horn for wide band millimeter-wave applications. IEEE Transactions on Antennas and Propagation
2004; 52(3):848854.
4. Yang D, Chung YC, Haupt R. Genetic algorithm optimization of a multisectional corrugated conical corn antenna.
Microwave and Optical Technical Letters 2003; 38(5):352356.
5. Dobson DC. Optimal mode coupling in simple planar waveguides. In IUTAM Symposium on Topological Design
Optimization of Structures, Machines, and Materials, Copenhagen, Denmark, Bense MP, Olhoff N, Sigmund O
(eds). Springer: Berlin, 2006; 311320.
6. Noreland D. Numerical techniques for acoustic modelling and design of brass wind instruments. Ph.D.
Thesis, Department of Information Technology, Uppsala University, Uppsala, Sweden, 2003. (Available from
http://publications.uu.se/theses/.)
7. Bangtsson E, Noreland D, Berggren M. Shape optimization of an acoustic horn. Computer Methods in Applied
Mechanics and Engineering 2003; 192:15331571.
8. Berggren M, Bangtsson E, Noreland D. Multifrequency shape optimization of an acoustic horn. In Computational
Fluid and Solid Mechanics 2003, Bathe KJ (ed.). Elsevier: Amsterdam, 2003; 22042207.
9. Jensen JS, Sigmund O. Systematic design of acoustic devices by topology optimization. Twelfth International
Congress on Sound and Vibration, ICVS12 2005, Lisbon, 2005.

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme
1606 R. UDAWALPOLA AND M. BERGGREN

10. Sigmund O, Jensen JS. Design of acoustic devices by topology optimization. In Short Papers of the 5th World
Congress on Structural and Multidisciplinary Optimization WC-SMO5, Venice, Italy, Cinquini C, Rovati M,
Venini P, Nascimbene R (eds). 2003; 267268.
11. Wadbro E, Berggren M. Topology optimization of an acoustic horn. Computer Methods in Applied Mechanics
and Engineering 2006; 196:420436.
12. Wadbro E, Berggren M. Topology optimization of wave transducers. In IUTAM Symposium on Topological Design
Optimization of Structures, Machines, and Materials, Copenhagen, Denmark, Bense MP, Olhoff N, Sigmund
O (eds). Springer: Berlin, 2006; 301310.
13. Engquist B, Majda A. Absorbing boundary conditions for the numerical simulation of waves. Mathematics of
Computation 1977; 31(139):629651.
14. Colton W, Kress R. Integral Equation Methods in Scattering Theory. Wiley: New York, 1983.
15. Wadbro E. On the far-field properties of an acoustic horn. Technical Report 2006-042, Department of Information
Technology, Uppsala University, Uppsala, 2006.

Copyright q 2007 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2008; 73:15711606
DOI: 10.1002/nme

S-ar putea să vă placă și