Sie sind auf Seite 1von 12

Acoust Aust (2016) 44:137–147

DOI 10.1007/s40857-016-0049-4

ORIGINAL PAPER

Boundary Element Modelling of a Novel Simple Enhanced


Bandwidth Schroeder Diffuser Offering Comparable Performance
to a Fractal Design
Andrew Lock1 · Damien Holloway1

Received: 21 December 2015 / Accepted: 16 February 2016 / Published online: 15 March 2016
© Australian Acoustical Society 2016

Abstract It is well established that Schroeder diffusers may be used as surface treatment to promote diffuse sound fields in
auditoria over a target frequency range. Quadratic Residue Diffusers are very effective within a design frequency bandwidth,
and this frequency bandwidth may be significantly extended by a nested or fractal design. Publicly available data typically
publish diffusion coefficients separately for the parent diffuser and the child diffuser, with some ambiguity surrounding the
effect the two levels have on each other, and the performance at ‘crossover’ frequencies. This paper uses an optimised 2D
boundary element method to investigate the differences in results when the parent and child diffusers are modelled separately,
and as an integrated design. It is found that the scattering properties of the child diffuser are significantly modified by the well
of the parent diffuser in which it is placed. More interestingly, the excellent scattering properties of the fractal design may
be achieved with only a minor loss of bandwidth using a far simpler and more robust geometry in which the child diffuser is
replaced by a concave surface at the base of the well in the parent diffuser. The paper outlines the optimisation of the boundary
element code for this task, and compares results for a variety of well bottom shapes with the fractal design.

Keywords Boundary element method · Diffusion · Fractal diffusers · Architectural acoustics

1 Introduction rooms and performance spaces, often to create a more even


sound field or to avoid strong reflections [1]. One distinct
An acoustic diffuser is an object or surface profile designed advantage of using diffusers as an acoustic tool is that they
to provide a diffuse reflection from an incident sound wave. generally possess low acoustic absorption properties, thereby
While nearly every reflective surface will offer some degree preserving sound energy and the reverberation characteristics
of diffusion, an acoustic diffuser is designed to produce high of a space. Where traditionally acoustic absorbers have been
levels of diffusion or specific diffusion characteristics, often used to avoid undesirable acoustic effects, there is a growing
for a design application or frequency range. Diffusion can be trend towards the use of diffusers instead [2], and this has
either temporal (time-related) or spatial, and is a useful tool increased the need to accurately model their performance.
in the acoustic treatment of spaces such as critical listening Schroeder diffusers, named after their creator Manfred
Schroeder [3], are popular due to their predictable and
This paper is based on the conference submission that received the desirable diffusion characteristics, and Quadratic Residue
Best Student Paper Award, sponsored by Resonate Acoustics, at the
Australian Acoustical Society Conference, November 15-18, 2015.
Diffusers (QRDs) are the most widely used and best-known
Schroeder diffuser type. While Schroeder diffusers offer
B Andrew Lock high-diffusion performance within a design bandwidth, this
andrew.john.lock@gmail.com bandwidth is limited by the practical dimensions of the
Damien Holloway diffuser and performance outside the bandwidth limits is
damien.holloway@utas.edu.au poor. Variations of ‘traditional’ Schroeder diffusers includ-
1 School of Engineering and ICT, University of Tasmania,
ing fractal designs (see Sect. 3.2) are used to increase this
Sandy Bay, TAS 7005, Australia design bandwidth while minimising the diffuser’s physical

123
138 Acoust Aust (2016) 44:137–147

dimensions [4]. However, the effectiveness of fractal diffuser ∇ 2  + k 2  = 0, (1)


designs with many thin panels and reverberant spaces may be
reduced due to higher absorption properties, therefore partly in terms of which particle velocity V and pressure p are,
offsetting the benefits of a complex diffuser when compared respectively,
to a simple acoustic absorber.
A new diffuser geometry is therefore proposed that
V = ∇, (2)
achieves a large bandwidth comparable to that of fractal dif-
∂
fusers, but with minimal additional complexity, through the p = −ρ = iρω. (3)
∂t
use of curved well bottoms.
The present study poses a challenge for numerical sim-
ulation since the length scales vary over two orders of Here ρ is the density of the acoustic medium, k = 2π λ =
ω
magnitude. This is a consequence of the wide frequency c is the wavenumber and λ, ω and c are, respectively, the
range studied and more particularly the fractal geometry wavelength, angular frequency and speed of sound.
employed. Thus an optimised Boundary Element Method Specific exterior problems are further specified by express-
was required. ing the total potential as the sum of an incident potential ϕ
The Boundary Element Method (BEM) is an algorithm (e.g. due to an unconfined sound source, typically a plane,
derived from Green’s Second Identity used for modelling and spherical or cylindrical wave) and a scattered (diffracted)
analysis of acoustic fields. The BEM is a powerful alterna- wave potential φ, both satisfying a radiation condition, and
tive to Finite Element Analysis (FEA), the Finite-Difference the sum of which satisfies the body surface boundary con-
Time Domain (FDTD) and other acoustic modelling tech- dition. For a rigid body, the latter condition is one of zero
niques, due to the nature of discretisation that effectively normal particle velocity, hence
reduces the dimensions of the problem by one. The efficient
nature of solving acoustic fields using the BEM is particularly ∇φ · n = −∇ϕ · n = −V I · n, (4)
suited to problems where the domain is large or considered
infinite, such as external diffusion of an object in an incident where V I is the incident wave local particle velocity, and n
sound field. is the unit normal to the body pointing into to the acoustic
This paper will firstly introduce optimised methods of domain (out of the body).
computing an external diffusion model with the BEM, with The BEM solves this by discretising Green’s second iden-
particular reference to Schroeder diffusers. The BEM is tity, derived from the divergence theorem, which may be
based on a program initially written by Rocchi [5], which expressed as
has been developed to achieve significant increases in speed
and accuracy through the implementation of advanced dis-
 
cretisation of the diffuser shape and efficient analytic and φ(q)∇G(p, q) · nq −G(p, q)∇φ(q) · nq dSq = −Cφ(p),
numeric integrals of terms used in matrix calculations.
The paper will then describe Schroeder diffusers and use (5)
this optimised BEM code to model performance of ‘tradi-
tional’ Schroeder diffusers, fractal diffuser designs and a where p is an arbitrary point within the acoustic domain and q
new diffuser design that offers diffusion performance com- is a point on the domain boundary and the integral is evaluated
parable with fractal designs but with a simpler, more robust over the entire domain boundary.
geometry expected to have favourable absorption proper- Of particular interest is when p is on the boundary, in
ties. which case a choice suitable for 2D exterior problems of
G(p, q) = 4i H0(1) (kr ) (where r = |q − p| and H0(1) (x) is the
Hankel function of the first kind and order zero [8]) leads
2 Boundary Element Method Development to C = 21 . By dividing the boundary (a closed curve for 2D
problems) into discrete small segments bounded by points qi ,
2.1 BEM Theory over which φ and ∇φ may for practical purposes be assumed
constant and equal to the value at the centroid, this becomes
Extended derivations of the Boundary Element Method a matrix equation
(BEM) may be found in several sources including [6] and
[7], a summary will be presented here. 1
− {φ} = [M]{φ} + [L]{V I · n}, (6)
Time periodic acoustic fields may be described in terms 2
of the real part of a velocity potential  = 0 e−iωt governed
by the Helmholtz equation where

123
Acoust Aust (2016) 44:137–147 139

 q  q
j j i (1) Table 1 Gauss integration points for various source-collocation dis-
Li j = G(pi , q) dsq = H0 (kr ) dsq , tances
q j−1 q j−1 4
Source-collocation distance r/λ Number of integration points N
(7)
 q  q
j −ik
j (1) <0.5 10
Mi j = ∇G(pi , q) · nq dsq = H1 (kr ) r · nq dsq ,
q j−1 q j−1 4 0.5–1 6
q − pi 1–3 4
r= . (8)
|q − pi | >3 2

The integrals L i j and Mi j are not generally known in closed


form and must be integrated numerically, which is very effi-
ciently accomplished using Gaussian quadrature [8]. The approximation is 2N − 1 where N is the number of integra-
self-influence terms (i = j) can however be evaluated ana- tion points (i.e. a 6-point rule is equivalent to fitting an 11th
lytically, and it is desirable to do so since they contain a order polynomial). Abscissae and weights for N = 2, 4, 6
singularity at r = 0. Some authors remove the singular lead- and 10 (amongst others) are published in Abramowitz and
ing term in the Taylor expansion (of order ln r or r1 ) and apply Stegun [8], and a good balance between efficiency and accu-
quadrature to the remainder, but using results in Abramowitz racy was found by adapting the integration rule depending
and Stegun [8] it can be shown on the normalised distance r of the source point q from the
collocation point p according to the rule in Table 1.
 a 
(1) (1) (1)
H0 (kr ) dr = 2a H0 (ka) + πa H0 (ka)H1 (ka) 2.2.2 Self-Influence Terms
−a
 |q −qi−1 |
(1)
− H1 (ka)H0 (ka) , a = i As already mentioned, G and ∇G have singularities at r = 0.
2
(9) Crude results are obtained using a quadrature rule with even
 a N , which does not include a point at r = 0 (unlike odd N ,
(1) 2 (1)
H1 (kr ) dr = − H0 (ka). (10) which cannot be evaluated at all). Substantially better results
−a k
are obtained however
 by numerically integrating
 the desingu-

(1) 2i (1) i ,
The Hankel function Hn(1) (a) is easily evaluated using a built- larised forms H0 (kr ) − π ln kr and H0 (kr ) + π2kr
in function in Matlab but the Struve function Hn (a) has to and adding the analytic integrals of 2i ln kr and −2i , which
π π kr
be evaluated using a truncated infinite series, thus are easily evaluated. But for relatively little additional com-
 putational cost (since this is only required for the diagonal
a 2  (1) ∞
H0(1) (kr ) dr = H0 (ka) sn elements) the full analytic integration presented in Eqs. (9)–
−a k n=0 (12) gave accuracy that was significantly better again.
∞
(1)
+ H1 (ka) tn , (11)
n=0
2.2.3 Code Restructure
where
The vast majority of run time is spent on computing the [L]
−(ka)2 and [M] matrices. Optimising the choice of quadrature rule
s0 = ka, sn = sn−1 ,
(2n − 1)(2n + 1) clearly therefore had a big impact on run time, but further
−(ka)2 significant speed gains were made by restructuring code to
t0 = (ka)2 , tn = tn−1 . (12) minimise recalculation of quantities and/or to exploit Mat-
(2n + 1)2
lab’s vector algebra capabilities. On the other hand some
These series converge rapidly with appropriate element Matlab functions, such as dot products of vectors with only
length relative to wavelength. two elements, could actually be performed faster by manu-
ally programming them.
2.2 BEM Optimisation Techniques
2.2.4 Treatment of Sharp Edges
2.2.1 Gaussian Quadrature Rules
A major source of errors is the steep pressure gradient in
As mentioned above, the integrals of Eqs. (7) and (8) are the vicinity of sharp edges (corners in 2D). Furthermore,
evaluated using Gaussian quadrature. Integration accuracy the coefficient C in Eq. (5) at the corner is not 1/2, but for
could be arbitrarily improved, at a cost, by increasing the an edge angle of θ is actually θ/2π (hence 1/2 for a smooth
order of the integration rule. For line integrals, the order of boundary, and it is noted that in the present BEM formulation

123
140 Acoust Aust (2016) 44:137–147

but the solution without the additional secondary domain


is perfectly suitable for these frequencies. An example of
an additional boundary within the diffuser body is shown in
Fig. 2.

2.2.6 Future Work

For the problem to be well conditioned, the element size on


any thin panels (such as those separating the diffuser wells)
had to be smaller than the panel thickness. This was quite
restrictive. Wu [9] has published a ‘thin panel’ BEM for-
mulation that overcomes this problem. This represents the
panel as a single surface so would not only allow larger ele-
Fig. 1 Surface pressure around an infinitely long 0.5 m × 0.5 m square ments, but would halve the number of elements on the panels
column due to a cylindrical source at 3240 Hz. The horizontal axis
(though with two equations for each).
represents distance around the perimeter, with edges at 0, 0.5, 1.0, 1.5
and 2.0 m. 0–0.5 m is the side facing the source Further gains could also be made by use of a higher order
formulation, in which φ and ∇φ may vary linearly on each
element, thus φ would be defined at element end points and
the boundary condition is applied at the element mid-point, would vary continuously around the boundary. This would
which is never a corner). Nevertheless this contributes to the be of benefit at corners in particular.
abrupt change of φ (or pressure) near the edge.
The steep pressure gradient at the edges is clearly visi-
ble in Fig. 1, showing a typical surface pressure around a 2.3 BEM Validation and Improvement Quantification
square object. It would not be a problem for shapes defined
by smooth curves, but is critical for modelling any object with Initial validation of the optimised BEM code was performed
edges, and similar results were found with the Schroeder dif- for the case of an infinite cylinder in a plane wave acoustic
fusers. field, for which there exist published solutions by Morse
A two-fold approach has been adopted to maintain accu- and Ingard [10], Kuyama [11] and Wiener [12]. The max-
racy. First, a 20-point quadrature rule was used on all imum deviation between a diffusion result from the BEM
elements attached to an edge. Second, the element size was and analytic solution published by Morse and Ingard [10]
refined approaching edges. Elements were reduced to 13 times was less than 0.003 % for a cylinder of radius a = k1 and
the global size at the corners, and transitioned in a geomet- observation radius r = 2k . The BEM code was further val-
ric progression over 5 elements, thus adjacent elements did idated for the more challenging case of a QRD with the
not differ by more than a ratio of 30.2 ∼= 1.25. A substantial effects of thin panels and sharp corners, including compar-
improvement in accuracy for a given number of elements was isons between results gained using alternative methods such
found from adopting both of these techniques, as discussed as the commercial Finite Element program ANSYS. Further
below. The authors plan to investigate this further in future information on the BEM code validation can be found in
work. Lock [7].
Figure 3 shows a comparison between the BEM code pub-
2.2.5 Singular Frequencies lished by Rocchi [5] and the BEM code developed in this
work. The BEM published by Rocchi [5] uses a uniform
At particular frequencies corresponding to the eigen values element size and computes the integrals of Eqs. (7) and (8)
of Eq. (6), the problem becomes poorly conditioned and using a 20-point Gaussian quadrature rule. The optimisations
accuracy substantially declines. This was solved by plac- described above produced more than 240 times increase in
ing an additional boundary inside the diffuser body, creating solution speed while maintaining or even improving accu-
a second acoustic domain entirely independent of the pri- racy. The lower half of Fig. 3 shows that for the test case at
mary domain. As this domain is not connected to the primary least 400 elements are required in the fully optimised code
domain, and all boundary and field conditions are enforced with edge refinement for reasonable convergence of results,
in both domains, it had no effect on the solution other than to requiring 0.25 min (15 s) to solve. Without edge refinement
change the eigen frequencies, hence to remove the singular- more than 1000 elements are required for the same accuracy,
ity at the problem frequencies and vastly improve accuracy. and the solution time is increased to more than 1 min. Without
Of course new problem frequencies are thus introduced, any optimisation 1000 elements took about 60 min.

123
Acoust Aust (2016) 44:137–147 141

Fig. 2 QRD geometry with


hollow cavity used to modify the
eigen frequencies of Eq. (6)

Fig. 3 Comparison of solution


time and convergence from 3
BEM code versions. Three
repeating periods of an N = 7
QRD, N wT = 0.56 m,
f = 2000 Hz = 4 f 0 (symbols
defined in Sect. 3.1)

3 Diffuser Design fusion lobes give Schroeder diffusers their high diffusion
properties.
3.1 Schroeder Diffusers In two dimensions, Schroeder phase grating diffusers con-
sist of wells of even width and varying depth separated by
Manfred Schroeder is considered to have made the most thin panels. Different mathematical patterns govern the depth
substantial breakthrough in diffuser design in 1975, when of successive wells, and installations most commonly consist
he presented his theory of reflection phase grating diffusers of multiple repeated periods of the chosen pattern. Although
[3,13]. Schroeder diffusers offered a key advantage in that some modern variations to Schroeder diffusers omit the thin
their performance could easily be predicted, and they are panels between wells in an effort to reduce absorption, they
still considered amongst some of the most effective dif- are retained during this work. A period consists of N wells
fuser designs used today. A key feature of phase grating of width w, separated by a thin panel of width wP . The
diffusers are regions of high sound pressure reflected in radial total width of a well and panel is represented by wT , and
directions from the diffuser, named ‘diffusion lobes’ most consequently the period width is written as N wT , shown in
prominent at multiples of the design frequency. These dif- Fig. 4.

123
142 Acoust Aust (2016) 44:137–147

Fig. 4 (Left) One period of a QRD; (Right) Repeating periods of a QRD array

c
3.1.1 Quadratic Residue Diffusers f max = . (16)
2w
Quadratic Residue Diffusers (QRDs) are named after the
pseudo-random quadratic residue sequence first studied by 3.2 High Bandwidth Diffusers
Gauss and Legendre [3], which dictates successive relative
well depths. The nth term in the sequence, sn , is defined as As shown in the previous section, each diffuser design has
an effective working range. In order to extend that range,
sn = n 2 modulo(N ), nested or fractal designs may be used [3]. Typically a ‘child’
QRD is placed in the bottom of each well of a parent QRD.
where n = 0, 1, 2 . . . (N − 1), and modulo (N ) represents For a diffuser of period N , the child QRD will be approx-
the least non-negative remainder when divided by N . The imately N times smaller than the parent (though this is not
value N is restricted to prime numbers, and also represents necessarily the case for the well depth). The performance
the number of terms in each repeating sequence. The QRD of the parent and child are generally published separately
nominal design frequency f 0 represents the lower limit of by manufacturers. However the child diffuser scatters sound
the design frequency bandwidth, and the depth dn of the nth within a confined space (the well), so its performance may
well in a quadratic residue diffuser for a prime number N is potentially differ significantly from its published unconfined
determined from the equation performance. Furthermore, there may be simpler geometries
csn that produce good scattering of high frequency sound within
dn = , (13)
2N f 0 a well.
It is of interest to determine what if any interaction there
where c is the speed of sound in air. is between parent and child that may enhance or degrade per-
formance, and to further investigate alternative geometries.
3.1.2 Design Frequency Limitations This paper therefore investigates the following four diffuser
designs:
A design principle behind Schroeder’s theory of phase grat-
ing diffusers is that longitudinal wave propagation within the
1. The Parent QRD is an N = 7 QRD of period width
wells dominates transverse waves. From this, the upper fre-
(N wT )parent = 0.85 m and design frequency f 0,parent =
quency limit to the applicability of this theory can be stated as
400 Hz.
λmin 2. The Child QRD is an N = 7 QRD approximately a 17
= w. (14) scale of the parent QRD, with (N wT )child = wparent =
2
10.93 cm and f 0,child = 7 f 0,parent = 2800 Hz.
Another limiting factor of QRDs is the period width, N wT . 3. The Fractal QRD is the Parent QRD with a Child QRD
While even energy lobes can be expected at the design fre- at the bottom of each well.
quency, f 0 , if the period width is too small only one energy 4. The Concave QRD is the Parent QRD with a concave
lobe is present in the specular reflection direction, and the surface at the bottom of each well. The concave curve
diffuser displays poor diffusion characteristics [2]. follows a quadratic profile with a maximum deviation of
The effective bandwidth of a simple QRD can therefore 1.35 cm = w8 .
be predicted as
csmax An array of 5 repeating periods was modelled for each dif-
f min = f 0 = , (15) fuser design, with the exception of the Child QRD for which
2N dmax
an array of 35 periods was modelled. Each diffuser design is
provided that N wT > λ0 , and shown in Fig. 5.

123
Acoust Aust (2016) 44:137–147 143

The reduced accuracy observed at incident angles can be


explained from the thin panel effect as described by [2] and
mentioned in Sect. 2.2.6. The implementation of the BEM
thin panel solution as published by [9] may resolve these
errors, but will result in a model where thin panels are consid-
ered to have infinitesimal thickness, and at high frequencies
the model may not accurately represent the physical diffuser.

4 Results
Fig. 5 Diagram of each diffuser modelled. 1 period of (a), (c), (d); 7
periods of (b) Figure 6 shows the normalised diffusion coefficient for the
four diffuser designs tested for the frequency range between
3.3 Test Details 100 Hz–10 kHz. Of particular interest is the comparative per-
formance of each diffuser design at the crossover frequencies
Each diffuser was modelled with a single cylindrical source between 1.5–4 kHz, and at high frequencies above 4 kHz. The
at normal (0◦ ) incidence at a distance of 20 m. Diffusion diffusion polar patterns for each diffuser at 2, 4 and 8 kHz
coefficients were based on sound pressures at 360 receivers are shown in Figs. 7, 8 and 9, respectively. At these frequen-
on a semicircle of 10 m radius, and 7 logarithmically spaced cies there are substantial differences in the performance of all
frequencies in each 1/3 octave bandwidth, in accordance with four diffuser designs, with the exception of 2 kHz where the
ISO 17497-2 [14]. The panel thickness relative to well width performance of the Fractal QRD, Concave QRD and Parent
wP
was w T
= 0.1. QRD are notably similar. Pressures shown in Figs. 7, 8 and 9
An accuracy study suggests the results in this paper are are scaled with respect to the maximum receiver pressure for
accurate to within ±2 % below 3 kHz, and within ±4 % each design.
between 3 and 10 kHz [7]. Accuracy was reduced with a Complex geometries, such as the Fractal QRD, required
source location at oblique angles to the diffuser, and there- over 5000 elements to obtain an accurate solution at high
fore, in this investigation, the model was limited to source frequencies, and obtaining the results in Fig. 6 required a
location at an angle θ = 0◦ to the centre of the diffuser. solution at 147 frequencies for each diffuser and a reference

Fig. 6 (Top) Normalised


diffusion coefficient of the
Parent, Child, and
corresponding Fractal QRD;
(Bottom) Normalised diffusion
coefficient of the Fractal QRD
and Concave QRD

123
144 Acoust Aust (2016) 44:137–147

Fig. 7 Polar plot example of the four tested diffuser designs at f = 2000 Hz. Normalised diffusion coefficients are a = 0.43, b = 0.002, c = 0.40
and d = 0.41

Fig. 8 Polar plot example of the four tested diffuser designs at f = 4000 Hz. Normalised diffusion coefficients are a = 0.15, b = 0.43, c = 0.65
and d = 0.51

123
Acoust Aust (2016) 44:137–147 145

Fig. 9 Polar plot example of the four tested diffuser designs at f = 8000 Hz. Normalised diffusion coefficients are a = 0.09, b = 0.50, c = 0.59
and d = 0.52

flat plate used to normalise the diffusion coefficient. For the varies for each frequency, whereas the specular component
case of a detailed diffuser such as a fractal QRD, a full set is present for all frequencies.
of results between 100 Hz–10 kHz required approximately Figure 7 also shows that at the lower boundary of the
12 h to compute. crossover frequency band there are minimal differences in
the scattering properties between the parent diffuser and con-
cave diffuser. This is consistent with the close agreement of
5 Discussion diffusion coefficients observed in Fig. 6 up until the crossover
frequency band, suggesting that introducing a nested concave
As shown in Fig. 6, it is observed that the fractal diffuser profile to well bottoms has little effect on the performance of a
offers slightly inferior performance when compared to the regular QRD within its design frequency band. It can be con-
corresponding parent diffuser at frequencies at which half cluded that below 2 kHz, the fine geometry in the fractal and
the wavelength is comparable to the well width, or the concave diffusers has no real effect and these will perform
‘crossover’ frequencies (i.e. around 1500 Hz). This is also very similarly to the parent diffuser (which will be similar to
evident from observation of the scattering polar patterns at the child diffuser at a frequency seven times higher).
2 kHz presented in Fig. 7. One explanation for this is the effect Figure 8 shows that at the top end of the crossover fre-
of the nested child diffuser on the quadratic residue sequence quency band the difference in scattering behaviour is much
interference pattern, resulting in reduced energy lobe behav- more profound than is suggested by the diffusion coefficient
iour. The small reduction in performance at the crossover alone. In particular, the reasonably good diffusion coefficient
frequencies is offset by significantly increased performance of the child diffuser hides the fact that there is significant lob-
at high frequencies, where the fractal diffuser offers far supe- ing. In contrast, both the fractal and concave diffusers show
rior diffusion to the parent or child diffuser when modelled far better diffusion, with relatively uniform pressures over
individually. The superior high-frequency performance and around 120◦ . Interestingly, the specular reflection peak evi-
extended bandwidth is consistent with previous findings by dent in the Fractal QRD is replaced with two individual peaks
Cox and D’Antonio [2]. The peak in the specular reflection at around ±10◦ in the Concave QRD.
direction in Fig. 7 may be attributed to the effect of aver- At 8 kHz (Fig. 9), the four diffusers all have quite differ-
aging 7 frequencies, where the angle of the diffusion lobes ent polar plots. The narrow quasi-specular pattern and low

123
146 Acoust Aust (2016) 44:137–147

diffusion coefficient of the parent diffuser shows that it is The diffusion performance of a fractal QRD was compared
clearly above its effective bandwidth. The strong lobing for to the corresponding parent QRD, child QRD and QRD with
the child diffuser contributes to its high diffusion coefficient, concave well bottoms. Key findings from the results include:
but the more uniform scattering of both the fractal and con-
cave diffusers (both also with higher diffusion coefficients) is • The performance of the Child QRD is modified, and gen-
more desirable. The fractal diffuser has slightly better spread- erally enhanced, by being incorporated within the Fractal
ing width, with significant energy over a 110◦ arc, but still QRD;
exhibits distinct lobing and a strong quasi-specular compo- • The Fractal QRD exhibited strong performance over a
nent. On the other hand the concave diffuser shows more wide frequency band, but there is a crossover frequency
uniformity of spreading, but only over about 75◦ , dropping band where performance is reduced by about 50 %;
off more rapidly at wider angles. Beyond ±65◦ from the nor- • A novel QRD design with concave well bottoms intro-
mal the fractal and concave diffusers are equally good and duced in this investigation has slightly different charac-
better than either parent or child. teristics at the crossover frequency band, but is generally
The novel concave diffuser design modelled in this work speaking comparable to the Fractal QRD design and
shows almost identical performance to the parent diffuser incorporates a far simpler and more robust design and
until the upper limit of the parent diffuser design bandwidth • There is scope to optimise the curve profile of the QRD
(around 2 kHz). At frequencies at which the wavelength is with concave well bottoms, which is likely to further
smaller than the well width, the concave diffuser design offers improve performance.
a level of performance comparable with the fractal diffuser
and superior to the child diffuser, and it is likely that this Acknowledgments The authors are grateful to Luca Rocchi, who
may be further improved through optimisation of the con- wrote the original BEM code that has been developed for this work.
cave curve profile. While the performance of the concave
diffuser may be considered slightly worse at mid-high fre-
quencies than the fractal diffuser, the simplicity of design
offers advantages including robustness, reduced production References
cost, and decreased risk of absorption. Fujiwara and Miya-
1. Everest, F.A., Pohlmann, K.C.: Master Handbook of Acoustics, 5th
jima recorded absorption coefficients between 0.3 and 1 for
edn. McGraw-Hill, New York (2009)
QRDs [15], which was later attributed to poor build quality 2. Cox, T., D’Antonio, P.: Acoustic Absorbers and Diffuser: Theory
and bonding of the tested diffusers [16]. The simplicity of the Design and Application. CRC Press, London (2009)
Concave QRD design in this work offers a clear advantage 3. Schroeder, M.R.: Diffuse sound reflection by maximum-length
sequences. J. Acoust. Soc. Am. 57(1), 149–150 (1975)
in reducing the risk of high absorption. While an estimate
4. D’Antonio, P., Konnert, J.: The QRD diffractal: a new one- or
of the absorption provided by diffusers can be predicted two-dimensional fractal sound diffusor. J. Audio Eng. Soc. 40(3),
numerically by applying a surface impedance to the diffuser 117–129 (1992)
surface, it is likely that ignoring effects such as structural 5. Rocchi, L.: Quadratic residue diffusers: modelling through the
boundary element method. BE(Hons) Thesis, University of Tas-
resonance in thin panels and bonded elements as well as
mania (2014)
absorption within viscous boundary layers will provide inac- 6. Kirkup, S.: The Boundary Element Method in Acoustics. Integrated
curate results. An experimental approach is recommended to Sound Software, Heptonstall (1998)
quantify any absorption benefits provided by the Concave 7. Lock, A.: Development of a 2D boundary element method to model
Schroeder acoustic diffusers, BE(Hons) Thesis University or Tas-
QRD design.
mania (2014)
8. Abramowitz, M., Stegun, I.A.: Handbook of Mathematical Func-
tions with Formulas, Graphs and Mathematical Tables. Applied
Mathematics Series, vol. 55. National Bureau of Standards, Wash-
ington, DC (1964)
6 Conclusion 9. Wu, T.W.: A direct boundary element method for acoustic radiation
and scattering from mixed regular and thin bodies. J. Acoust. Soc.
This investigation developed an optimised implementation Am. 97(1), 84–91 (1995)
of the BEM to model external diffusion from Schroeder 10. Morse, P.M., Ingard, K.U.: Theoretical Acoustics. Princeton Uni-
versity Press, Princeton (1986)
diffusers. Increases in speed and accuracy gained from 11. Kuyama, T.: Some calculated results of the diffraction of sound
advanced discretisation of the diffuser shape and analytic waves by a cylindrical obstacle. Proc. Physico-Mathematical Soc.
integration of diagonal matrix elements, amongst other devel- Jpn. 15, 366–370 (1933)
opments, allowed for modelling of geometries with greater 12. Wiener, F.M.: Sound diffraction by rigid spheres and circular cylin-
ders. J. Acoust. Soc. Am. 19(3), 444–451 (1947)
levels of detail including large arrays of small QRDs, fractal 13. Schroeder, M.R.: Binaural dissimilarity and optimum ceilings for
QRDs and a novel QRD design with concave well bot- concert halls: more lateral sound diffusion. J. Acoust. Soc. Am.
toms. 65(4), 958–963 (1979)

123
Acoust Aust (2016) 44:137–147 147

14. ISO 17497-2:2012: Acoustics—Sound-Scattering Properties of 16. Fujiwara, K.: A study on the sound absorption of a quadratic-
Surfaces Part 2: Measurement of the Directional Diffusion Coef- residue type diffuser. Acta Acust. United Acust. 81(4), 370–378
ficient in a Free Field. International Organization for Standardiza- (1995)
tion, Geneva (2012)
15. Fujiwara, K., Miyajima, T.: Absorption characteristics of a prac-
tically constructed Schroeder diffuser of quadratic-residue type.
Appl. Acoust. 35(2), 149–152 (1992)

123
Acoustics Australia is a copyright of Springer, 2016. All Rights Reserved.

Das könnte Ihnen auch gefallen