Sound Reproduction Systems Using Variable-Directivity Loudspeakers

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Sound reproduction systems using variable-directivity loudspeakers

M. A. Poletti, F. M. Fazi, and P. A. Nelson

Citation: The Journal of the Acoustical Society of America 129, 1429 (2011); doi: 10.1121/1.3533689
View online: https://doi.org/10.1121/1.3533689
View Table of Contents: https://asa.scitation.org/toc/jas/129/3
Published by the Acoustical Society of America

ARTICLES YOU MAY BE INTERESTED IN

Sound-field reproduction systems using fixed-directivity loudspeakers


The Journal of the Acoustical Society of America 127, 3590 (2010); https://doi.org/10.1121/1.3409486

Interior and exterior sound field control using general two-dimensional first-order sources
The Journal of the Acoustical Society of America 129, 234 (2011); https://doi.org/10.1121/1.3518772

Analysis and control of multi-zone sound field reproduction using modal-domain approach
The Journal of the Acoustical Society of America 140, 2134 (2016); https://doi.org/10.1121/1.4963084

Generation of an acoustically bright zone with an illuminated region using multiple sources
The Journal of the Acoustical Society of America 111, 1695 (2002); https://doi.org/10.1121/1.1456926

Theory and design of sound field reproduction in reverberant rooms


The Journal of the Acoustical Society of America 117, 2100 (2005); https://doi.org/10.1121/1.1863032

Acoustic contrast, planarity and robustness of sound zone methods using a circular loudspeaker array
The Journal of the Acoustical Society of America 135, 1929 (2014); https://doi.org/10.1121/1.4866442
Sound reproduction systems using variable-directivity
loudspeakers
M. A. Polettia)
Industrial Research Ltd., P.O. Box 31-310, Lower Hutt, New Zealand

F. M. Fazi and P. A. Nelson


Institute of Sound and Vibration Research, University of Southampton, United Kingdom
(Received 22 April 2010; revised 23 November 2010; accepted 8 December 2010)
Sound reproduction systems using omnidirectional loudspeakers produce reflections from room surfaces
which interfere with the desired sound field within the array. While active compensation systems
can reduce the reverberant level, they require calibration in each room and are processor-intensive.
Directional loudspeakers allow the direct to reverberant level to be improved within the array, but still
produce a finite exterior field which reflects from the room surfaces. The use of variable-directivity
loudspeakers allows the exterior field to be eliminated at low frequencies by implementing the
Kirchhoff–Helmholtz integral equation. This paper investigates the performance of variable-directivity
arrays in reducing reverberant levels and compares the results with those derived in a previous paper for
fixed-directivity arrays. The results presented may have some impact on the design of commercial
multi-channel systems for sound reproduction.
C 2011 Acoustical Society of America. [DOI: 10.1121/1.3533689]
V

PACS number(s): 43.60.Tj, 43.55.Jz, 43.38.Md, 43.60.Sx [WMC] Pages: 1429–1438

I. INTRODUCTION in a volume of space with boundary S using monopole sour-


ces and dipole sources normal to the surface S and that the
Commercially available surround sound systems approxi-
sound field outside that region is zero. The dipole and
mate two-dimensional (2D) sound reproduction using a rela-
monopole amplitudes differ for each loudspeaker, and hence
tively small number of loudspeakers in a circular array around
the K–H integral can be implemented using first-order vari-
a listener.1,2 The use of a greater number of loudspeakers
able-directivity loudspeakers.
allows the holographic reproduction of wave fields over a
This paper considers sound reproduction using variable-
larger volume of space using the 2D circular or three-dimen-
directivity approaches based on the K–H integral. A spherical
sional (3D) spherical arrays.3–17 The loudspeakers used in the
loudspeaker array is assumed, and so the loudspeaker first-
current systems are typically omnidirectional at low frequencies
order responses are a weighted combination of a monopole
and therefore, when used in typical listening rooms, produce
and a radially oriented dipole response. We show that a
early reflections and reverberation which add to the desired
mode-truncated solution produces superior results to a solu-
direct field produced within the array. The effects of the room
tion based on the direct discretization of the K–H integral,33
can be reduced by acoustical treatment or by using active com-
and we also develop a mode-matching solution and compare
pensation systems which measure various directional responses
it with the mode-truncated solution. Finally, we evaluate vari-
at the desired listener position and then pre-compensate the
able-directivity solutions by numerical simulations and com-
loudspeaker signals to eliminate the room effects.18–25
pare the results with the fixed-directivity results derived in
A simpler reduction of room effects is possible by using
Ref. 26.
directional loudspeakers which increase the level of direct
The theory in this paper follows in an analogous manner
sound relative to the reverberant field.26,27 In Ref. 26, the
to that presented in Ref. 26 and we therefore briefly review
use of fixed-directivity loudspeakers in 3D sound reproduc-
the relevant results from Ref. 26 and then extend it to the
tion systems has been studied, and it has been shown that the
variable-directivity case.
use of hyper-cardioid loudspeakers improves the direct to
reverberant sound ratio by a factor approaching 4, which
accords with the characteristics of a single hyper-cardioid II. SPHERICAL HARMONIC DESCRIPTION OF SOUND
FIELDS
loudspeaker.28,29
A greater reduction of room effects is possible using an Consider a sphere XrL of radius rL centered on the ori-
array of variable-directivity loudspeakers. These can be con- gin. A sound field is referred to as an interior field if it satis-
figured to reduce the exterior sound from the array using the fies the homogeneous wave equation in the interior of XrL
Kirchhoff–Helmholtz (K–H) integral formula.9–13,30–32 The and as an exterior field if it satisfies the homogeneous wave
K–H integral states that a desired sound field can be created equation in the exterior of XrL .30
The solution of the wave equation can be expressed in
spherical coordinates ~r ¼ ðr; h; /Þ, where the arrow denotes
a)
Author to whom correspondence should be addressed. Electronic mail: a vector quantity, the vector norm r ¼ k~ r k is the radial dis-
m.poletti@irl.cri.nz tance from the origin, h is the elevation angle from the

J. Acoust. Soc. Am. 129 (3), March 2011 0001-4966/2011/129(3)/1429/10/$30.00 C 2011 Acoustical Society of America
V 1429
sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
vertical z-axis, and / is the azimuthal angle from the x- ð2n þ 1Þ ðn  jmjÞ! jmj
axis.30 Assuming a harmonic time dependence of exp(ixt), Ynm ðh; /Þ ¼ P ðcos hÞeim/ ; (2)
4p ðn þ jmjÞ! n
the spherical harmonic expansions of an interior and exterior
sound field at angular frequency x are
where Pjnmj ðÞ is the associated Legendre polynomial. The
8 1 n assumption is made that the operating frequency x and
> X X
>
> jn ðkrÞAm m
>
< n ðk ÞYn ðh; /Þ; r < rL hence the wave number k are fixed, and therefore the de-
pðr; h; /; kÞ ¼
n¼0 m¼n pendence of the sound field and other functions of x or k is
>X
>
1 X n
not written explicitly.
>
> hn ðkr ÞCm m
: n ðk ÞYn ðh; /Þ; r > rL ;
n¼0 m¼n
A. Monopole and dipole sources
(1)
The acoustic pressure field generated by an ideal
where k ¼ x=c is the wave number, c is the speed of sound monopole source in the free field is of the form30
in meters per second (assumed to be uniform in R3 Þ, Am n ðk Þ
and Cmn ð k Þ are the expansion coefficients, j n(x) is the nth eikj~r~rs j
pm ð~ rs Þ ¼ Gð~
r;~ r j~
rs Þ ¼ ; (3)
order spherical Bessel function, and hn ð xÞ ¼ hðn1Þ ð xÞ is the r ~
4pj~ rs j
nth order spherical Hankel function of the first kind. The
spherical harmonic Ynm ðh; /Þ is defined as32 which has a spherical harmonic expansion

8
> X
1 X
n
>
> ik j ð kr Þh ð kr Þ Ynm ðh; /ÞYnm ðhs ; /s Þ ; r < rs
>
< n n s
n¼0 m¼n
pm ð~ rs Þ ¼
r;~ (4)
>
> X1 Xn
>
> jn ðkrs Þhn ðkrÞ Ynm ðh; /ÞYnm ðhs ; /s Þ ; r > rs ;
: ik
n¼0 m¼n

where r ¼ k~ r k and rs ¼ k~ rs k. first-order directivity that is approximately independent of


A dipole at position ~ rs and oriented in direction ~ v has a frequency. At a given distance rd from the dipole source, the
field that takes the form31 equalized response is flat for frequencies down to the transi-
  tion frequency fd ¼ c=(2prd) and rises at 6 dB per octave
@Gð~ r j~
rs Þ eikj~r~rs j i below that. Hence, to ensure flat responses down to a fre-
pd ð~ rs Þ ¼
r;~ ¼ ik 1þ cos c;
@~m r ~
4pj~ rs j r ~
kj~ rs j quency fd, we must maintain a distance from any equalized
(5) dipole greater than rd.
The spherical harmonic expansion for a 1/(ik)-equalized
where c is the angle between ~ r ~
v and ~ rs . As in Ref. 26, we dipole oriented radially with respect to the origin is obtained
equalize the dipole response by dividing by ik to produce a from the derivative of Eq. (4) with respect to rs

8 1 n
> X X
>
> k jn ðkrÞh0n ðkrs ÞYnm ðh; /ÞYnm ðhs ; /s Þ ; r < rs
>
<
n¼0 m¼n
pd ðr; h; /Þ ¼ (6)
>
> X1 X n
>
> j0n ðkrs Þhn ðkr ÞYnm ðh; /ÞYnm ðhs ; /s Þ ; r > rs ;
:k
n¼0 m¼n

where j0n ðÞ and h0n ðÞ are the derivatives of the corresponding are presented. For an evaluation of the exterior sound field pro-
spherical Bessel and Hankel functions. duced by the array the exterior truncation error is also useful.
The angle-averaged normalized truncation error for a
monopole with order N expansion PT(r, h, /) is defined as6,8
B. Interior and exterior truncation error of monopole
and dipole fields ð
If the series in Eqs. (4) and (6) are truncated to a given fi- jpðr; h; /Þ  pT ðr; h; /Þj2 dXr
nite order, n ¼ N, the representation of the field is no longer eTM ðN; kr Þ ¼ Xr ð ; (7)
exact since it is affected by the so-called truncation error. In jpðr; h; /Þj2 dXr
Ref. 26, the interior truncation error of the monopole and dipole Xr

1430 J. Acoust. Soc. Am., Vol. 129, No. 3, March 2011 Poletti et al.: Variable-directivity speaker sound systems
lower error is required, particularly for kr close to krs, then
N > krs is required.
The truncation error for the dipole is found from Eq. (6)
8
> XN  0 
>
> ð 2n þ 1Þj 2
ð kr Þ h ðkrs Þ2
>
> n n
>
>
>
>
n¼0
1X ; kr < krs
>
> 1  0 2
>
> 2 
ð2n þ 1Þjn ðkrÞ hn ðkrs Þ 
>
<
n¼0
eTD ðN; kr Þ ¼
>
> XN
>
> ð2n þ 1Þj02 2
>
> n ðkrs Þjhn ðkr Þj
>
>
> n¼0
>1  X
>
> 1 ; kr > krs :
>
> ð 2n þ 1 Þj02
ð kr Þ j h ð kr Þ j 2
: n s n
n¼0
(9)

FIG. 1. Monopole truncation error for krs ¼ 8 and orders N ¼ 0 to 10. This is shown in Fig. 2 for krs ¼ 8. The error is around
16 dB for N  kr and kr  krs and for kr > krs, the error
where the overbar denotes an average over all angles. Substi- reduces asymptotically to around 12 dB for N  krs, 2 dB
tuting the monopole expansions, Eq. (4), yields the monop- higher than the monopole case.
ole truncation error In summary, to represent accurately the exterior field
due to a general first-order source, we require N > krs. For
8 the interior representation of the source, N ¼ kr is sufficient,
> XN
>
> ð2n þ 1Þj2n ðkrÞjhn ðkrs Þj2 implying that low orders are sufficient for small radii.26
>
>
>
> Finally, for better accuracy near the source, orders higher
>
>
n¼0
1  X ; kr < krs
>
> 1
than krs are required.
>
> ð 2n þ 1Þj 2
ð kr Þ j h ð kr Þ j2
>
< n n s
n¼0
eTM ðN; kr Þ ¼
>
> XN
C. Loudspeaker array geometry and Nyquist
>
> ð2n þ 1Þj2n ðkrs Þjhn ðkr Þj2
>
> frequencies
>
>
>
>
n¼0
1  X ; kr > krs :
>
> 1
We will assume a spherical array of loudspeakers, each
>
> 2
ð2n þ 1Þjn ðkrs Þjhn ðkr Þj 2
: of which can produce ideal monopole and dipole fields,
n¼0
arranged at positions (rL, h1, /1), l [ [1, L], which approxi-
(8)
mate a uniform sampling over the sphere, with numerical
weighting coefficients bl, required for accurate numerical
The truncation error is shown for krs ¼ 8 in Fig. 1. For
integration. A measure of the uniformity of the loudspeaker
kr  krs the error is approximately 14 dB for N  kr.6 For
arrangement is given by the discrete cross-correlation of the
kr > krs, the minimum order required to represent the exte-
spherical harmonics at the loudspeaker angles
rior expansion at large kr is N  krs, where the asymptotic
error again falls to around 14 dB or lower. However, if
X
L
Cððn1 ; m1 Þ; ðn2 ; m2 ÞÞ ¼ bl Ynm11 ðhl ; /l ÞYnm22 ðhl ; /l Þ : (10)
l¼1

This matrix indicates the degree of orthonormality of the


sampled spherical harmonics for the transducer arrangement
adopted. In the ideal case of perfect orthonormality, the
cross-correlation equals dn1n2 dm1m2, where dnm is the Kro-
necker delta.26 In practice at high orders, the sampled spher-
ical harmonics are not orthogonal and this will contribute to
reproduction errors at high frequencies.26 The array is able
to reproduce a sound field at a given wave number k and at
radius r for N > kr and hence [since there are (N þ 1)2
spherical harmonics of order N] if L  ðdkr e þ 1Þ2 uni-
formly arranged loudspeakers are used (see Ref. 6). This
defines approximately the interior Nyquist frequency of the
array26
pffiffiffi 
L1
c
fNI ðrÞ ¼ : (11)
FIG. 2. Dipole truncation error for krs ¼ 8 and orders N ¼ 0 to 10. 2pr

J. Acoust. Soc. Am., Vol. 129, No. 3, March 2011 Poletti et al.: Variable-directivity speaker sound systems 1431
For the control of the exterior field for r > rL Fig. 2 shows scaled by rL2 bl to produce discrete monopole weights ul and
that we require L  ðdkrL e þ 1Þ2 . Therefore, surround sys- dipole weights vl.
tems which attempt to control the exterior field are able to
do so up to the fixed exterior Nyquist frequency B. Mode-truncated solution
pffiffiffi  Equations (1), (4), and (6) have spherical harmonic expan-
c L1
fNE ¼ : (12) sions of infinite order. When implementing the integral in Eq.
2prL (13) using a discrete array as in Eq. (16), aliasing will occur for
those spherical harmonics which cannot be controlled with the
Hence, the interior Nyquist frequency increases toward the ori- loudspeaker array.33 This aliasing can be reduced by using
gin due to the focusing of the individual sound fields and the mode-truncated forms of the monopole and dipole amplitudes.33
exterior Nyquist frequency is fixed since the Nyquist constraint In view of Eqs. (1) and (13), the order-N truncated expansion
ensures that the correct exterior field is generated near the array. for the monopole amplitude in Eq. (16) is
III. SOUND REPRODUCTION SYSTEMS WITH X
N X
n
VARIABLE-DIRECTIVITY LOUDSPEAKERS u^l ¼ krL2 bl j0n ðkrL ÞAm m
n ðk ÞYn ðhl ; /l Þ: (17)
n¼0 m¼n
We now consider the implementation of the K–H inte-
gral. As in Ref. 26, we assume a continuous distribution of
Including the minus sign in the dipole weight amplitudes,
monopole and radially oriented dipole speakers on the sur-
the mode-truncated expansion of the dipole amplitude is
face of a sphere XrL of radius rL at positions ~ rv ¼ ðrL ; h v ;
derived from Eqs. (1) and (13)
/v Þ, hv 2 ½0; p, and /v 2 ½0; 2p. The exterior field is zero,
and we write the interior field expression as X
N X
n

ð   v^l ¼ ikrL2 bl jn ðkrL ÞAm m


n ðkÞYn ðhl ; /l Þ: (18)
@pð~
rv Þ 1 @Gð~rj~
rv Þ n¼0 m¼n
pð~
rÞ ¼ r j~
Gð~ rv Þ  ikpð~
rv Þ dXrL ;
X rL nð~
@~ rv Þ nð~
ik @~ rv Þ
Substituting these solutions into Eq. (16) shows that, below
r < rL ; (13) the Nyquist frequency, they produce the correct sound field
Eq. (1) if the loudspeaker positions produce the ideal ortho-
where pð~ rv Þ is the sound pressure on the boundary of the normality conditions in Eq. (10).
region XrL , dXrL ¼ rL2 sinðhV Þ dhV d/V is an element of solid For the case of a point source, we obtain the following
angle, and where we have explicitly included the ik equaliza- expressions for the mode-truncated weights:
tion of the dipole.
X
N X
n
A. Direct solution for a single point source u^l  iðkrL Þ2 bl j0n ðkrL Þhn ðkrs ÞYnm ðhs ; /s Þ Ynm ðhl ; /l Þ
n¼0 m¼n
For the case where the sound field is due to a point
source outside the region at ~
rs , the monopole source ampli- (19)
tudes are, from Eq. (5),
and
rs ~
ikj~ rj
 
@pð~
rÞ e i
¼ ik 1þ cos t; (14) X
N X
n
@n rs  ~
4pj~ rj rs  ~
kj~ rj v^l  ðkrL Þ2 bl jn ðkrL Þhn ðkrs ÞYnm ðhs ; /s Þ Ynm ðhl ; /l Þ:
n¼0 m¼n
where m is the angle between the normal and rs. The dipole (20)
source amplitudes are, from Eq. (3),

eikj~rs ~rj C. Mode-matching solution


ikpð~
rÞ ¼ ik : (15)
rs  ~
4pj~ rj As for the fixed-directivity case,26 a mode-matching so-
lution can be derived for variable-directivity sound repro-
In practice a discrete array is used, using L loudspeakers duction systems. This allows some control of the effects of
rl ¼ ðrL ; hl ; /l Þ on XRL . The sound field
arranged at positions ~ non-orthogonality of the array at high mode orders. We
p^ð~
rÞ reproduced by a discrete array approximation of determine the loudspeaker weights required to synthesize the
Eq. (13) can be represented by field due to a monopole source, while requiring the exterior
L   sound field to be zero. For simplicity, we assume that the
X r j~
1 @Gð~ rl Þ
p^ð~
rÞ ¼ r j~
Gð~ rl Þ ul þ vl ; (16) monopole source generating the target field is outside the
l¼1
ik @~ rl Þ
nð~ loudspeaker array. We require the interior spherical har-
monic expansions of the sum of sound fields due to the L
where ul and vl are the monopole and equalized dipole loudspeakers, with monopole weights ul [applied to Eq. (4)
amplitudes, respectively. for position ~rl ¼ ðrL ; hl ; /l Þ] and radial dipole weights vl ,
As for the fixed-directivity case, the weights in Eq. (14) [applied to Eq. (6)] to equal the desired field expansion in
and (15) are calculated for each loudspeaker position and Eq. (1). For each (n, m), this produces

1432 J. Acoust. Soc. Am., Vol. 129, No. 3, March 2011 Poletti et al.: Variable-directivity speaker sound systems
X
L  terms Y ¼ Ynm ðhl ; /l Þ . Since U is block diagonal, its inverse
k vl Ynm ðhl ; /l Þ ¼ Am
ul þ h0n ðkrL Þ
ihn ðkrL Þ n; is a block diagonal matrix of 2 2 inverse matrices
l¼1
 1  1  0 
n 2 ½0; N ; m 2 ½n; n: (21) 1 iH H0 D 0 J H0
U ¼ ¼ ; (27)
iJ J0 0 D1 iJ iH
Similarly, since we require zero exterior field, the exterior
mode-matching equations for radii greater than the loud- where D is the diagonal determinant D ¼ i[HJ0 – JH0 ], which
speaker radius r > rL are has the Wronskian property30
1
X
L  D ¼ i½HJ0  JH0  ¼ I (28)
k vl Ynm ðhl ; /l Þ ¼ 0;
ul þ j0n ðkrL Þ
ijn ðkrL Þ ðkrL Þ2
l¼1
and hence
n 2 ½0; N ; m 2 ½n; n : (22)
 
1 2 J0 H0
We note that Eq. (17) and (18) satisfy these two equations U ¼ ðkrL Þ : (29)
iJ iH
assuming discrete orthonormality of the spherical harmonics,
Cððn1 ; m1 Þ; ðn2 ; m2 ÞÞ ¼ dn1 ;n2 dm1 ;nm [Eq. (10)]. Therefore, the mode-matching matrix in Eq. (23) can be written
Equations (21) and (22) can be written in matrix     0    0 
notation Y 0 u J H0 d Jd
¼ krL2 ¼ krL2 : (30)
0 Y v iJ iH 0 iJd
      
WIM WID u u d
¼W ¼ ; (23) The mode-matching solution can therefore be written in the
WEM WED v v 0
alternative form
8  
where WIM is a K ¼ (N þ1 )2 by L matrix of interior monop- >  H 1 H J0 d
>
>
  < L 2 2kr 2
Y Y þ k Y ; KL
ole terms WIM ðb; lÞ ¼ ikhn ðkrL ÞYnm ðhl ; /l Þ , where b ¼ n2 þ n u
V 2
iJd
þ m þ 1, WID is a K L matrix of interior dipole terms ¼   (31)
v >
>  1 J0 d
WID ðb; lÞ ¼ kh0n ðkrL ÞYnm ðhl ; /l Þ , u is an L 1 vector of > 2 H H
: krL Y2 Y2 Y2 þ kV ; K < L:
monopole weights ul , and v is an L 1 vector of dipole iJd
weights vl . Similarly, WEM is a K L matrix of exterior Furthermore, the mode-truncated solutions in Eq. (17) and
monopole terms WEM ðb; lÞ ¼ ikjn ðkrL ÞYnm ðhl ; /l Þ and WED is (18) can be written in matrix form as
a K L matrix of exterior dipole terms WED ðb; lÞ ¼ kj0n ðkrL Þ    H  0   0 
Ynm ðhl ; /l Þ . Finally, d is a K 1 vector of desired sound field ^
u 2 Y 0 Jd 2 H Jd
¼ krL B ¼ krL BY2 ; (32)
terms db ¼ Am n and 0 is a K 1 vector of zeros. ^v 0 YH iJd iJd
For K  L , the least squares solution to Eq. (23) is35
where B is an L L diagonal matrix of weighting coefficients
    bl. It is now clear from a comparison of Eq. (31) with Eq. (32)
u  H 1 H d
¼ W W þ kV I W ; (24) that the mode-matching solution contains an additional inverse
v 0
matrix which compensates for the non-uniformity of the loud-
where kV is a regularization parameter and the superscript H speaker array geometry, whereas the mode-truncated solution
denotes the conjugate transpose. For K < L, the minimum assumes orthogonality of the sampled spherical harmonics in
energy solution is36 using the Hermitian transpose of Y2 to obtain the associated
solution.8
   
u H
 H
1 d
¼ W WW þ kV I : (25) E. Robustness
v 0
As discussed in Ref. 26, the sensitivity of the loud-
speaker array to variations in the loudspeaker polar responses
D. Comparison of solutions and errors in the array layout is approximately governed by
The matrix W in Eq. (23) can be decomposed into the the condition number of the K L matrix WV.34 This matrix
product of 2 2 block matrices has a singular value decomposition WV ¼ USK VH, where U
is a K K unitary matrix, SK is a K L matrix containing (at
   full rank) K singular values [r1, r2, … , rK], and V is an L L
iH H0 Y 0
W¼k ¼ kUY2 ; (26) unitary matrix.34 For K < L, the K squared singular values of
iJ J0 0 Y
WV are the eigenvalues of
where H and H0 are the diagonal matrices whose terms are W V WH 2 H H H H
V ¼ USK U ¼ UY2 Y2 U  UU ; (33)
spherical Hankel functions and their first derivative, respec-
tively, and analogously J and J0 are diagonal matrices with where the last term assumes that the matrix Y2 is close to
spherical Bessel functions and their derivative, respectively. unitary. Therefore, the condition number is
Note that each of these matrices have 2nþ1 repeated terms


for each value of n. Y is a K L matrix of spherical harmonic jV ¼ kWV k


W1 V

¼ kUk
U1
; (34)

J. Acoust. Soc. Am., Vol. 129, No. 3, March 2011 Poletti et al.: Variable-directivity speaker sound systems 1433
IV. SIMULATIONS
We now present simulations of sound fields produced
by variable-directivity arrays in free-field conditions using
the mode-truncated and mode-matched solutions. For com-
parison with the fixed-directivity case,26 we also present
simulations using arrays with weights given by
X
N X
n
Am m
 n Yn ðhl ; /l Þ ;
^l ¼ bl
w 0
l 2 ½1; L;
n¼0 m¼n k aihn ðkrL Þ þ ð1  aÞhn ðkrL Þ
(39)

where a is a first-order weighting parameter which produces


monopole loudspeaker responses for a ¼ 1 and dipole responses
for a ¼ 0. With a ¼ 0.25, each loudspeaker polar response is a
hyper-cardioid which produces the maximum direct to rever-
berant ratio in reverberant conditions.
We use a spherical array of 144 loudspeakers which are
FIG. 3. Log(j) at four frequencies from 100 to 600 Hz, K ¼ (Nþ1)2 < L
approximation and random normal matrix result, L ¼ 144, rL ¼ 1.5 m.
approximately uniformly arranged34 at a radius of 1.5 m
producing an exterior Nyquist frequency of 400 Hz. The
near-field of the dipole responses occur at a radius r from the
where ||
|| is the two-norm, equal to the largest singular value origin for which k (rL  r) ¼ 1, which is a radius of 1.22 m at
of WV. The norm ||U1|| is found from Eq. (29) 200 Hz. Due to the poor conditioning of the mode-matching

 0 
for the maximum possible mode order of 11 (see Fig. 3) and

1

H0
the need to maintain a high order for exterior control, we

U
¼ ðkrL Þ2
J
¼ ðkrL Þ2 kUk; (35)

iJ iH
will use N ¼ 10 in the simulations. In this case, the mode-
matched solution is given by Eq. (25) and the approximate
since the adjugate matrix adj{U} has the same singular val- condition number in Eq. (38) is valid.
ues as U. Hence The desired field is that due to a point source positioned
on the x-axis at rs ¼ 3 m. The error performance was found

jV ¼ ðkrL Þ2 kUk2 ¼ ðkrL Þ2 max eig UUH : (36) to be similar for all source angles, due to the regular arrange-
ment of the loudspeakers, and the fact that the source dis-
Since U is block diagonal, the eigenvalues of UUH are sim- tance is larger than the loudspeaker radius (so that the source
ply the (repeated) eigenvalues of the set of 2 2 matrices position never coincides with a loudspeaker position).
U2 UH2 where
A. Reproduction error and exterior field level
 
ihn ðkrL Þ h0n ðkrL Þ To quantify the performance of the solutions inside the
U2 ðn; krL Þ ¼ ; n ¼ 0; N: (37)
ijn ðkrL Þ j0n ðkrL Þ array, we calculate the angle-averaged (radial) relative error
between the desired field p(r, h, /, k) and the reproduced
The largest eigenvalues occur for the maximum order n ¼ N field p^ðr; h; /; kÞ at radius r < rL (Ref. 6)
and so ð
jpðr; h; /; kÞ  p^ðr; h; /; kÞj2 dXr
jV  ðkrL Þ2 maxfeigfU2 ðN; krL ÞU2 ðN; krL ÞH gg: (38)
eðkrÞ ¼ Xr ð : (40)
The condition number of W is shown in Fig. 3 for four fre- jpðr; h; /; kÞj2 dXr
Xr
quencies from 100 to 600 Hz. Equation (38) correctly pre-
dicts the condition number for K < L, when K ¼ L there is an This may be determined for a point source with spherical

increase in condition number, and Eq. (38) is not correct for harmonic expansion coefficients Am m
n ¼ ihn ðkrs ÞYn ðhs ; /s Þ
K > L. At 600 Hz (above the Nyquist frequency), the condi- using the orthogonality properties of the spherical harmonic
tioning becomes similar to that of the random matrix case, expansions in Eqs. (1), (4), and (6) yielding (using as an
with a peak of E{log(jV)} ¼ log(2L) þ 0.982 ¼ 6.64.37 example the mode-matching coefficients ul and vl )

 2
X
1 Xn XL   m 
 0  m 
4p j2n ðkrÞ  i ul hn ðkrL Þ þ vl hn ðkrL Þ Yn ðhl ; /l Þ  ihn ðkrs ÞYn ðhs ; /s Þ 
n¼0

m¼n l¼1

eðkrÞ ¼ : (41)
X1
2
2
ð2n þ 1Þjn ðkr Þjhn ðkrs Þj
n¼0

1434 J. Acoust. Soc. Am., Vol. 129, No. 3, March 2011 Poletti et al.: Variable-directivity speaker sound systems
To determine the performance of the surround system outside the array we calculate the angle-averaged exterior sound pres-
sure magnitude squared, relative to the angle-averaged desired sound pressure magnitude squared at the loudspeaker radius
ð
jp^ðr; h; /; kÞj2 dXr
 ðkrÞ ¼ ð Xr
! ; r > rL (42)
jpðrL ; h; /; kÞj2 dXrL
X rL

which can be written in terms of the spherical harmonics for a point source field in the form

 2
X
1 X n X L   
 2  
jhn ðkrÞj  i ul jn ðkrL Þ þ vl j0n ðkrL Þ Ynm ðhl ; /l Þ 
n¼0

m¼n l¼1


!ðkrÞ ¼ 4p : (43)
X
1
ð2n þ 1Þj2n ðkrL Þjhn ðkrs Þj2
n¼0

Equations (41) and (43) may also be used with the mode- Figures 5 and 6 show the real part of the sound fields (at
truncated K–H weights u^l and v^l in Eqs. (19) and (20). In t ¼ 0) obtained from the truncated-fixed solution with
practice, we use a finite maximum order of n ¼ 25 in Eqs. a ¼ 0.25 (hyper-cardioid) and the variable-directivity solu-
(41) and (43) which produced negligible truncation error. tion [Eqs. (19) and (20)]. (The corresponding mode-matched
solution wave fields were similar in appearance, and their
B. Simulation results relative error performance is considered shortly.) The dashed
circle indicates the loudspeaker
pffiffiffi radius
 and the dark circle is
We first demonstrate the improvement in accuracy
the maximum radius rN ¼ L  1 =k where the mode-lim-
obtained using the mode-truncated form of the K–H integral
ited reproduction can maintain accuracy. The sound field
Eqs. (19) and (20) compared to the direct form Eqs. (14) and
produced by the fixed-directivity solution propagates across
(15) (corresponding to infinite order expansion).33 Figure 4
the interior of the array and beyond it. In contrast, the energy
shows the interior error and exterior relative field for both
of the variable-directivity field is largely confined to the inte-
solutions at 200 and 600 Hz. At low frequencies, the direct
rior of the array and the exterior field is very small. This is
and mode-truncated results are essentially the same. At 600
shown more clearly by the radial error, shown in Fig. 7.
Hz, the amplitudes of the high-order terms for N > 10 are
Mode-truncated and mode-matched results are shown for
significant, hence the error is larger for the direct solution,
both fixed- and variable-directivity cases, together with the
with an error of 17 dB at r ¼ 0, compared to less than
80 dB for the truncated case. The exterior error is around
5 dB higher for the direct solution. We, therefore, consider
only the mode-truncated and mode-matched solutions in the
following simulations.

FIG. 4. Interior radial error (r < 1.5 m) and relative exterior sound level FIG. 5. (Color online) Fixed-directivity sound fields for f ¼ 200 Hz, rL ¼ 1.5
(r > 1.5 m) for direct and mode-truncated K–H solutions at 200 and 600 Hz. m, rs ¼ 3 m, and a ¼ 0.25.

J. Acoust. Soc. Am., Vol. 129, No. 3, March 2011 Poletti et al.: Variable-directivity speaker sound systems 1435
solution errors reduce with radius and are below 80 dB at
the center of the array.
The mode-matched solutions are able to provide a more
consistent accuracy across a wider area than the mode-trun-
cated solutions and can reduce the error at larger radii.
Increasing the regularization raises the error floor (by around
20 dB in this example).
The variable-directivity relative exterior sound level
falls to below 70 dB at r ¼ 3 m for the mode-matched solu-
tion with kV ¼ 0.0001, to below 50 dB for kV ¼ 0.001, and
to around 65 dB for the mode-truncated case. These results
are over 40 dB lower than the fixed-directivity exterior levels
which both reduce to around 10 dB. Hence at low frequen-
cies (below the exterior Nyquist frequency), the exterior field
is significantly reduced by the variable-directivity array
compared to the fixed-directivity array.
The truncation error for the monopole source is also
shown in Fig. 7. It is below 80 dB for radii less than 1 m.
The reproduced sound fields do not approach this lower limit
since all positions greater than r ¼ 0.27 m are in the near-
field of the loudspeakers, where the truncation error for the
FIG. 6. (Color online) Variable-directivity sound fields for f ¼ 200 Hz, loudspeaker sources is large (Figs. 1 and 2).
rL ¼ 1.5 m, and rs ¼ 3 m. Figures 8 and 9 show the fixed-directivity and variable-
directivity mode-truncated sound fields produced at 600 Hz,
interior truncation error for the source [Eq. (8)], which repre- above the exterior Nyquist frequency of the array (400 Hz).
sents the lowest possible error for a truncation limit of In both cases, the interior field is accurate to a radius rN ¼ 1
N ¼ 10. The mode-matching errors are shown for regulariza- m. Both arrays now produce an appreciable exterior field.
tion parameters of kV ¼ 0.0001 and 0.001 to show how they The variable-directivity field is smaller in the direction to-
vary with the regularization. ward the source (at x ¼ 3 m) and on the downstream side of
The interior errors are approximately the same for both the array from the source but produces larger radiation in lat-
the fixed- and the variable-directivity solutions, showing that eral directions. The fixed-directivity solution does not reduce
there is no significant penalty associated with the exterior the downstream propagation, and the sound wave propagates
cancellation of the sound field. The mode-matching errors out of the array on the far side. There is also some lateral
are approximately constant with radius near the center of the radiation of sound producing interference effects.
array. They are around 50 dB for r < 1m for kV ¼ 0.001
and are about 70 dB for kV ¼ 0.0001. The mode-truncated

FIG. 7. Variable mode-truncated and mode-matched radial error (r < 1.5


m), and relative exterior sound level (r > 1.5 m) for f ¼ 200 Hz, a ¼ 0.25,
rL ¼ 1.5 m, rs ¼ 3 m, and N ¼ 10, for regularization parameters of 0.001 and
0.0001. The fixed-directivity mode-truncated and mode-matched results and FIG. 8. (Color online) Fixed-directivity mode-truncated sound field for
truncation error for N ¼ 10 and r < 1.5 m are also shown for comparison. f ¼ 600 Hz, rL ¼ 1.5 m, r ¼ 3 m, and a ¼ 0.25.

1436 J. Acoust. Soc. Am., Vol. 129, No. 3, March 2011 Poletti et al.: Variable-directivity speaker sound systems
the array to its exterior. The fixed-directivity solution pro-
duces a simpler exterior field in the region opposite to the vir-
tual source location, but a more significant backward radiation
of sound toward the virtual source. Generally the fixed-direc-
tivity exterior sound level is slightly lower than the variable-
directivity level, and so at frequencies above the exterior
Nyquist frequency, a fixed-directivity solution is preferable.

V. CONCLUSIONS
This paper has investigated the production of three-
dimensional sound fields using loudspeakers with variable-
directivity with the goal of reducing the low-frequency
reverberant field that occurs with sound reproduction in
rooms. The exterior—and hence the reverberant—field can
be eliminated below the exterior Nyquist frequency of the
array with no appreciable increase in interior reconstruction
error in comparison with fixed-directivity solutions. Above
this frequency, the array cannot cancel the exterior field and
fixed-directivity is more effective.
This paper has demonstrated that low-frequency control
of room acoustics is possible in 3D surround systems without
FIG. 9. (Color online) Variable-directivity mode-truncated sound fields for
f ¼ 600 Hz, rL ¼ 1.5 m, and rs ¼ 3 m.
the need to take into account the modal behavior of the
room. The sound system does not need to be calibrated in-
situ and is therefore simpler than adaptive systems, but it
The angle-averaged radial error is shown in Fig. 10. The
does require first-order loudspeakers with variable-directiv-
interior sound fields produce an error close to the truncation
ity which have closely matched polar responses. Further-
limit, but the mode-match solutions are able to maintain ac-
more, it does not eliminate sound scattered from the listener
curacy down to r ¼ 0.5 m, whereas the mode-truncated solu-
and reflected from the room surfaces.
tions diverge from the minimum error at r ¼ 0.8 m. The
We also note that the theory presented here is applicable
exterior field levels produced by the variable-directivity sol-
to any practical situation where 3D wave fields are synthe-
utions are now 2 dB higher than the fixed-directivity level
sized using an open sphere of directional transducers and
and so the benefit of variable-directivity is lost.
that such an array avoids the modal resonances that would
At higher frequencies, the variable-directivity solution
be produced by an array inside a solid enclosure (if the loud-
tends to produce radiation lobes at a variety of angles on the
speakers were arranged on a rigid wall). However, it does
downstream side of the array, caused by the unsuccessful
produce finite exterior fields above the exterior Nyquist fre-
attempt to cancel the sound propagating from the interior of
quency of the array.
The results in this paper are restricted to numerical
simulations and we have not investigated practical variable-
directivity loudspeakers, or the phase and amplitude match-
ing performance of typical transducers. Since exterior
cancellation is only possible at low frequencies, a practical
loudspeaker product would provide monopole and dipole
responses for the woofer (either as a number of fixed combi-
nations for fixed-directivity applications or with discrete
inputs for variable-directivity use) and the tweeter would
produce a single response with the directivity governed by
the loudspeaker baffle and tweeter diaphragm size. Such
loudspeakers would also prove beneficial in stereo sound
systems for reducing the effect of low-frequency modes. An
alternative approach would be to use separate woofer and
tweeter arrays, with a larger number of tweeters to extend
the accurate region of the interior sound field and reduce the
exterior field at higher frequencies.

FIG. 10. Variable mode-truncated and mode-matched radial errors (r < 1.5
ACKNOWLEDGMENTS
m) and relative exterior field (r > 1.5 m) for f ¼ 600 Hz, a ¼ 0.25, rL ¼ 1.5
m, rs ¼ 3 m, and N ¼ 10. The fixed-directivity mode-truncated and mode-
This work has been partially funded by the Royal Acad-
matched results and truncation error for N ¼ 10 and r < 1.5 m are also shown emy of Engineering and by the Engineering and Physical
for comparison. Sciences Research Council.

J. Acoust. Soc. Am., Vol. 129, No. 3, March 2011 Poletti et al.: Variable-directivity speaker sound systems 1437
1 20
M. A. Gerzon, “Ambisonics in multichannel broadcasting and video,” J. T. Betlehemxy and T. Abhayapala, “Theory and design of sound field
Audio Eng. Soc. 33(11), 859–871 (1985). reproduction in reverberant rooms,” J. Acoust. Soc. Am. 117, 2100–2111
2
M. A. Gerzon, “Optimum reproduction matrices for multispeaker stereo,” (2005).
21
J. Audio Eng. Soc. 40(7/8), 571–589 (1992). P.-A. Gauthier and A. Berry, “Adaptive wave field synthesis for sound
3
M. Gerzon, “Periphony: With-height sound reproduction,” J. Audio Eng. field reproduction: Theory, experiments, and future perspectives,” J.
Soc. 21(1), 2–10 (1973). Audio Eng. Soc. 55(12), 1107–1124 (2007).
4 22
O. Kirkeby and P. A. Nelson, “Reproduction of plane wave sound fields,” P.-A. Gauthier, A. Berry, and W. Woszczyk, “Sound-field reproduction
J. Acoust. Soc. Am. 94(5), 2992–3000 (1993). in-room using optimal control techniques: Simulations in the frequency
5
M. A. Poletti, “A unified theory of horizontal holographic sound systems,” domain,” J. Acoust. Soc. Am. 117(2), 662–678 (2005).
23
J. Audio Eng. Soc. 48(12), 1155–1182 (2000). S. Spors, M. Renk, and R. Rabenstein, “Limiting effects of active room
6
D. B. Ward and T. D. Abhayapala, “Reproduction of a plane-wave sound compensation using wave field synthesis,” in Proceedings of the 118th
field using an array of loudspeakers,” IEEE Trans. Speech, Audio Process. Convention of AES, Barcelona, Spain (May 28–31, 2005).
24
9(6), 697–707 (2001). S. Spors, H. Buchner, R. Rabenstein, and W. Herbordt, “Active listening
7
J. Daniel, “Spatial sound encoding including near field effect: Introducing room compensation for massive multichannel sound reproduction systems
distance coding filters and a viable new ambisonics format,” in Proceed- using wave-domain adaptive filtering,” J. Acoust. Soc. Am. 122(1), 354–
ings of the 23rd International Conference of AES (2003). 369 (2007).
8 25
M. A. Poletti, “Three-dimensional surround sound systems based on L. Fuster, J. J. Lopez, A. Gonzalez, and P. D. Zuccarello, “Room
spherical harmonics,” J. Audio Eng. Soc. 53(11), 1004–1025 (2005). compensation using multichannel inverse filters for wave field synthesis
9
A. J. Berkhout, D. de Vries, and P. Vogel, “Acoustic control by wave field systems,” in Proceedings of the 118th Convention of AES (May 28–31,
synthesis,” J. Acoust. Soc. Am. 93(5), 2764–2778 (1993). 2005).
10 26
M. M. Boone, E. N. G. Verheijen, and P. F. Van Tol, “Spatial sound-field M. A. Poletti, F. M. Fazi, and P. A. Nelson, “Sound-field reproduction sys-
reproduction by wave-field synthesis,” J. Audio Eng. Soc. 43(12), 1003– tems using fixed-directivity loudspeakers,” J. Acoust. Soc. Am. 127(6),
1012 (1995). 3590–3601 (2010).
11 27
D. de Vries, “Sound reinforcement by wavefield synthesis: Adaptation of M. M. Boone and. O. J. Ouweltjes, “Design of a loudspeaker system with
the synthesis operator to the loudspeaker directivity characteristics,” J. a low-frequency cardiodlike radiation pattern,” J. Audio Eng. Soc. 45(9),
Audio Eng. Soc. 44(12), 1120–1131 (1996). 702–707 (1997).
12 28
S. Takane, Y. Suzuki, and T. Sone, “A new method for global sound field W. Anhert and F. Steffen, Sound reinforcement engineering (E and FN
reproduction based on Kirchhoff’s integral equation,” Acust. Acta Acust. Spon, London, 1999), pp. 79–125.
29
85, 250–257 (1999). R. P. Glover, “A review of cardioid type unidirectional microphones,”
13
S. Ise, “A principle of sound field control based on the Kirchhoff- J. Acoust. Soc. Am. 11(3), 296–302 (1940).
30
Helmholtz integral equation and the theory of inverse systems,” Acust. E. G. Williams, Fourier Acoustics (Academic Press, San Deigo, 1999),
Acta Acust. 85, 78–87 (1999). pp. 183–234, 251–295.
14 31
N. Epain and E. Friot, “Active control of sound inside a sphere via control E. Skudrzyk, The Foundations of Acoustics (Springer-Verlag, New York, 1971),
of the acoustic pressure at the boundary surface,” J. Sound Vib. 299, 587– pp. 344–377.
32
604 (2007). D. Colton and R. Kress, “Inverse Acoustic and Electromagnetic Scattering
15
F. M. Fazi, P. A. Nelson, J. E. N. Christensen, and J. Seo, “Surround sys- Theory,” in Applied Mathematical Sciences, 2nd ed. (Springer, Berlin,
tem based on three-dimensional sound field reconstruction,” in Proceed- 1998), Vol. 93, pp. 21–28.
33
ings of the 125th Convention of AES (October 2–5, 2008). S. Spors and J. Ahrens, “A comparison of wave field synthesis and
16
F. M. Fazi and P. A. Nelson, “A theoretical study of sound field recon- higher-order ambisonics with respect to physical properties and spatial
struction techniques,” in Proceedings of the 19th International Congress sampling,” in Proceedings of the 125th Convention of AES (October 2–5,
on Acoustics (September 2–7, 2007). 2008).
17 34
J. Ahrens and S. Spors, “An analytical approach to sound field reproduc- http://www.personal.soton.ac.uk/jf1w07/nodes/nodes.html, last accessed March
tion using circular and spherical loudspeaker distributions,” Acta Acust. 3, 2010.
35
Acust. 94, 988–999 (2008). G. H. Golub and C. F. Van Loan, Matrix Computations (Johns Hopkins
18
P. A. Nelson, F. Orduna-Bustamante, and H. Hamada, “Multi-channel sig- University Press, Baltimore, MD, 1983) pp. 136–188.
36
nal processing techniques in the reproduction of sound,” J. Audio Eng. L. L. Scharf, Statistical Signal Processing: Detection, Estimation and
Soc. 44(11), 973–989 (1996). Time Series Analysis (Addison-Wesley Publ. Co., Massachusetts, 1991),
19
P. A. Nelson, H. Hamada, and S. J. Elliot, “Adaptive inverse filters for pp. 359–422.
37
stereophonic sound reproduction,” IEEE Trans. Signal Process. 40(7), A. Edelman, “Eigenvalues and condition numbers of random matrices,”
1621–1632 (1992). SIAM J. Matrix Anal., Vol. 9, No. 4, 543–560 (October 1988).

1438 J. Acoust. Soc. Am., Vol. 129, No. 3, March 2011 Poletti et al.: Variable-directivity speaker sound systems

You might also like