Download as pdf or txt
Download as pdf or txt
You are on page 1of 34

aeroacoustics volume 2 á number 2 á 2003 Ð pages 95 Ð 128 95

Surface integral methods in


computational aeroacousticsÑFrom the
(CFD) near-field to
the (Acoustic) far-field *
Anastasios S. Lyrintzis 
School of Aeronautics and Astronautics
Purdue University
W. Lafayette, IN 47907-2023

ABSTRACT
A review of recent advances in the use of surface integral methods in Computational
AeroAcoustics (CAA) for the extension of near-field CFD results to the acoustic far-field is
given. These integral formulations (i.e. Kirchhoff’s method, permeable (porous) surface Ffowcs-
Williams Hawkings (FW-H) equation) allow the radiating sound to be evaluated based on
quantities on an arbitrary control surface if the wave equation is assumed outside. Thus only
surface integrals are needed for the calculation of the far-field sound, instead of the volume
integrals required by the traditional acoustic analogy method (i.e. Lighthill, rigid body FW-H
equation). A numerical CFD method is used for the evaluation of the flow-field solution in the
near field and thus on the control surface. Diffusion and dispersion errors associated with wave
propagation in the far-field are avoided. The surface integrals and the first derivatives needed can
be easily evaluated from the near-field CFD data. Both methods can be extended in order to
include refraction effects outside the control surface. The methods have been applied to helicopter
noise, jet noise, propeller noise, ducted fan noise, etc. A simple set of portable Kirchhoff/FW-H
subroutines can be developed to calculate the far-field noise from inputs supplied by any
aerodynamic near/mid-field CFD code.

1 BACKGROUNDÑAEROACOUSTIC METHODS
For an airplane or a helicopter, aerodynamic noise generated from fluids is usually very
important. There are many kinds of aerodynamic noise including turbine jet noise,
impulsive noise due to unsteady flow around wings and rotors, broadband noise due to
inflow turbulence and boundary layer separated flow, etc. (e.g. Lighthill1). Accurate
prediction of noise mechanisms is essential in order to be able to control or modify

*Presented at the CEAS Workshop “From CFD to CAA” Athens, Greece, Nov. 2002.
†Professor, e-mail: lyrintzi@ecn.purdue.edu.
96 Surface integral methods in computational aeroacoustics

them to comply with noise regulations, i.e. Federal Aviation Regulations (FAR) part 36,
and achieve noise reductions. Both theoretical and experimental studies are being
conducted to understand the basic noise mechanisms. Flight-test or wind-tunnel test
programs can be used, but in either case difficulties are encounted such as high expense,
safety risks, and atmospheric variability, as well as reflection problems for wind-tunnel
tests. As the available computational power increases numerical techniques are
becoming more and more appealing. Although complete noise models have not yet
been developed, numerical simulations with a proper model are increasingly being
employed for the prediction of aerodynamic noise because they are low-cost and
efficient. This research has led to the emergence of a relatively new field: Computational
AeroAcoustics (CAA).
CAA is concerned with the prediction of the aerodynamic sound source and the
transmission of the generated sound starting from the time-dependent governing
equations. The full, time-dependent, compressible Navier-Stokes equations describe
these phenomena. Although recent advances in Computational Fluid Dynamics (CFD)
and in computer technology have made first-principle CAA plausible, direct extension
of current CFD technology to CAA requires addressing several technical difficulties
in the prediction of both the sound generation and its transmission.2–3 A review of
aerospace application of CAA methods was given by Long et al.4
Aerodynamically generated sound is governed by a nonlinear process. One class of
problems is turbulence generated noise (e.g. jet noise). An accurate turbulence model is
usually needed in this case. A second class of problems involves impulsive noise due to
moving surfaces (e.g. helicopter rotor noise, propeller noise, fan noise etc.). In these
cases an Euler/Navier-Stokes model or even a full potential model is adequate, because
turbulence is not important.
Once the sound source is predicted, several approaches can be used to describe
its propagation. The obvious strategy is to extend the computational domain for the
full, nonlinear Navier-Stokes equations far enough to encompass the location where
the sound is to be calculated. However, if the objective is to calculate the far-field
sound, this direct approach requires prohibitive computer storage and leads to
unrealistic turnaround time. The impracticality of straight CFD calculations for
supersonic jet aeroacoustics was pointed out by Mankbadi et al.5 Furthermore, because
the acoustic fluctuations are usually quite small (about three orders of magnitude less
than the flow fluctuations), the use of nonlinear equations (whether Navier-Stokes or
Euler) could result in errors, as pointed out by Stoker and Smith.6 One usually has no
choice but to separate the computation into two domains, one describing the nonlinear
generation of sound, the other describing the propagation of sound. There are several
alternatives to describing the sound propagation once the source has been identified.

1.1 Field solution of simpler equations


Linearized Euler Equations (LEE) The first alternative is to use simpler equations
in the acoustic far-field. The Linearized Euler Equations (LEE) have been used in order
to extend the CFD solutions to the far-field (e.g. Lim et al.7, Viswanathan and Sankar8,
Shih et al.9). The LEE equations employ a division of the flow field into a
aeroacoustics volume 2 á number 2 á 2003 97

time-averaged flow and a time-dependent disturbance which is assumed to be small.


The hybrid (zonal) approach consists of the near-field evaluation using an accurate
CFD code (e.g. for jet noise the code is usually based on Large Eddy Simulations: LES)
and the extension of the solution to the mid-field using LEE. Considerable CPU savings
can be realized, since the LEE calculations are much cheaper than the CFD calculations.
This approach is very promising, because it accounts for a variable sound velocity
outside the near-field where usually an LES model is applied. This method may also be
appropriate for an intermediate region in some problems, outside from the reactive
near-field where the speed of sound is still not constant, before moving to another
integral method for the far-field.
Other Equations Hardin and Pope10 have proposed a decoupling of the time-
dependent incompressible flow and the compressible aspects (acoustics) of the flow.
This technique was used successfully to predict the flow over a two-dimensional cavity.
A field solution of the wave equation can also be used (e.g. Freund11). Freund11 claims
that the field solution of the wave equation is cheaper than the surface integral solutions
(see section 1.2.2), when the solution everywhere in the field is sought. However, in
most applications only a few locations are needed to study directivity and compare with
microphone measurements. Also, for any numerical solution of field equations
dissipation and dispersion errors still exist and an accurate description of propagating
far-field waves is compromised.

1.2 Integral methods


1.2.1 Volume integral methods
Traditional Acoustic Analogy The first integral approach for acoustic propagation
is the acoustic analogy.12 In the acoustic analogy, the governing Navier-Stokes
equations are rearranged to be in wave-type form. There is some question as to which
terms should be identified as part of the sound source and retained in the right-hand
side of the equation and which terms should be in the left-hand side as part of the
operator (e.g., Lilley13). The far-field sound pressure is then given in terms of a volume
integral over the domain containing the sound source. Several modifications to
Lighthill’s original theory have been proposed to account for the sound-flow
interaction or other effects. The major difficulty with the acoustic analogy, however, is
that the sound source is not compact in supersonic flows. Errors could be encountered
in calculating the sound field, unless the computational domain could be extended in
the downstream direction beyond the location where the sound source has completely
decayed. Furthermore, an accurate account of the retarded time-effect requires keeping
a long record of the time-history of the converged solution of the sound source, which
again represents a storage problem. The Ffowcs Williams and Hawkings (FW-H)
equation14 was introduced to extend acoustic analogy in the case of solid surfaces.
However, when acoustic sources (i.e., quadrupoles) are present in the flowfield a
volume integration is needed. This volume integration of the quadrupole source term
is difficult to compute and is usually neglected in most acoustic analogy codes
(e.g. WOPWOP 15). Recently, there have been some successful attempts in evaluating
this term (e.g. WOPWOP+ 16,17).
98 Surface integral methods in computational aeroacoustics

1.2.2 Surface integral methods


Kirchhoff Method Another alternative is the Kirchhoff method which assumes that
the sound transmission is governed by the simple wave equation. Kirchhoff’s method
consists of the calculation of the nonlinear near- and mid-field, usually numerically,
with the far-field solutions found from a linear Kirchhoff formulation evaluated on a
control surface surrounding the nonlinear-field. The control surface is assumed to
enclose all the nonlinear flow effects and noise sources. The sound pressure can be
obtained in terms of a surface integral of the surface pressure and its normal and time
derivatives. This approach has the potential to overcome some of the difficulties
associated with the traditional acoustic analogy approach. The method is simple and
accurate and accounts for the nonlinear quadrupole noise in the far-field. Full
diffraction and focusing effects are included while eliminating the propagation of the
reactive near-field.
This idea of matching between a nonlinear aerodynamic near-field and a linear
acoustic far-field was first proposed by Hawkings.18 The separation of the problem into
linear and nonlinear regions allows the use of the most appropriate numerical
methodology for each. The terminology “Kirchhoff method” was introduced by George
and Lyrintzis.19 It has been used to study various aeroacoustic problems, such as
propeller noise, high-speed compressibility noise, blade-vortex interactions, jet noise,
ducted fan noise, etc. The use of Kirchhoff’s method has increased substantially the last
10 years, because of the development of reliable CFD methods that can be used for the
evaluation of the near-field. An earlier review on the use of Kirchhoff’s method was
given by Lyrintzis.20
Porous FW-H equation A final alternative is the use of permeable (porous)
surface FW-H equation. The usual practice is to assume that the FW-H integration
surface corresponds to a solid body and is impenetrable. However, if the surface is
assumed to be porous, a general equation can be derived (as shown in the original
reference 14 and in reference 21). The porous surface can be used as a control surface
in a similar fashion as the Kirchhoff method explained above. Thus the pressure signal
in the far-field can be found based on quantities on the control surface provided by a
CFD code.
Farassat in a recent review article22 reviewed all the available FW-H and Kirchhoff
equations for application to noise evaluation from rotating blades. The current article
focuses only on control surface methods (i.e. Kirchhoff, porous FW-H) and discusses
issues with their application in various types of aerocoustic problems including rotor
noise, jet noise, ducted fan noise, airfoil noise etc.). At first the main formulations will
be reviewed, advantages and disadvantages of each method will be discussed. Then we
will present several algorithmic issues and various application examples.

2 SURFACE INTEGRAL FORMULATIONS


2.1 KirchhoffÕs method formulations
Kirchhoff’s method is an innovative approach to noise problems which takes advantage
of the mathematical similarities between the aeroacoustic and electrodynamic
equations. The considerable body of theoretical knowledge regarding electrodynamic
aeroacoustics volume 2 á number 2 á 2003 99

field solutions can be utilized to arrive at the solution of difficult noise problems.
Kirchhoff’s formula was first published in 1882.23 It is an integral representation
(i.e. surface integral around a control surface) of the solution to the wave equation.
Kirchhoff’s formula, although primarily used in the theory of diffraction of light and in
other electromagnetic problems, it has also many applications in studies of acoustic
wave propagation.
The classical Kirchhoff formulation is limited to a stationary surface. Morgans24
derived a formula for a moving control surface using Green’s functions. Generalized
functions can also be used for the derivation of an extended Kirchhoff formulation.
A field function is defined to be identical to the real flow quantity outside a control
surface S and zero inside. The discontinuities of the field function across the control
surface S are taken as acoustic sources, represented by generalized functions. Ffowcs-
Williams and Hawkings14 derived an extended Kirchhoff formulation for sound
generation from a vibrating surface in arbitrary motion. However, in their formulation
the partial derivatives were taken with respect to the observation coordinates and time
and that is difficult to use in numerical computations. Farassat and Myers25 derived
a Kirchhoff formulation for a moving, deformable, piecewise smooth surface. The same
partial derivatives were taken with respect to the source coordinates and time. Thus their
formulation is easier to use in numerical computations and their relatively simple
derivation shows the power of generalized function analysis.
It should be noted that Morino and his co-workers26–30 have developed several
formulations for boundary element methods using the Green’s function approach. These
are based on the solution of the wave equation and hence, the integral expressions are
the same as in Kirchhoff’s method. However, the formulation in terms of the velocity
potential. This has advantages (e.g., the boundary condition is simple) as well as
disadvantages ( e.g., the pressure of the wake). Morino’s formulations were derived with
aerodynamic applications in mind, so the observer is in the moving coordinate system.
However, they can be used for aeroacoustics, for example when both the control surface
and the observer move with a constant speed (e.g., wind tunnel experiments), as
mentioned in reference 20. Their latest formulation29 appears to provide an integrated
boundary element framework for Aerodynamics and Aeroacoustics. A detailed
discussion about the differences in the aerodynamics and the aeroacoustics of their various
formulations can be found in reference 30.

2.1.1 FarassatÕs formulation


Farassat’s Kirchhoff formulation gives the far-field signal, due to sources contained
within the Kirchhoff surface. Assume the linear, homogeneous wave equation,

2 1 ¶ 2f ¶ 2f
f = 2 2
- =0 (1)
ao ¶ t ¶x i¶xi
is valid for some acoustic variable f , and sound speed ao, in the entire region outside
of a closed and bounded smooth surface, S .
The signal, in the stationary coordinate system, is evaluated with a surface integral
over the control surface, S , of the dependent variable, its normal derivative, and its time
100 Surface integral methods in computational aeroacoustics

derivative (Figure 1). S is allowed to move in an arbitrary rigid-body fashion. The


dependent variable f is normally taken to be the disturbance pressure, but can be any
quantity which satisfies the linear wave equation.

r é E1 ù é f E2 ù
4pf ( x , t ) = ò
S
ê ú dS +
êë r (1 - M r ) úû ret ò S
ê 2 ú dS
êë r (1 - Mr ) úû ret
(2)

where

(
E1 = Mn2 - 1 ) ¶¶fn + M n M t × Ñ 2f -
Mn Ç
ao
f+
1
ao (1 - M r ) 2
[
MÇ r ( cos q - Mn )f ]
+
1
ao (1 - Mr )
[
( nÇ r - MÇ n - nÇ M ) f + ( cos q - Mn ) fÇ + (cos q - Mn ) f ] (3)

(1 - M 2 )
E2 = ( cos q - Mn ) (4)
(1 - Mr ) 2
r r
Here ( x , t ) are the observer coordinates and time, and ( y, t ) are the source (surface)
coordinates and time. Mi is the Mach number vector of the surface, r is the distance
from source to observer, q is the source emission angle, and n̂ is the control surface unit
normal vector (cos q = rˆ × nˆ ). M t is the Mach number vector tangent to the surface, and
Ñ 2 is the surface gradient operator. A dot indicates a source time derivative, with the
position on the surface kept fixed. Also,

MÇ r = MÇ i rˆi nÇ r = nÇ i rˆi MÇ n = MÇ i nˆi nÇ M = nÇ i Mi (5)

Observer (x, t)

n
r

(y, t¢)

Control surface S
Figure 1. Kirchhoff’s surface S and notation.
aeroacoustics volume 2 á number 2 á 2003 101

The form of Equation (2) and E1 , E2 were given by Farassat and Myers.25 E2 was
presented in the simplified form shown here by Myers and Hausmann.31 The surface
integrals are over the control surface S , subscript ret indicates evaluation of the
integrands at the emission (retarded) time, which is the root of
r r
|x-y|
g=t -t+ =0 (6)
ao
If the frame velocity is subsonic at the surface, then equation (6) has a unique solution.
However, equation (2) is still valid for supersonically moving surfaces. As we can see
from equations 2 through 5, the (1 - Mr ) term can produce a singularity in the case
where the Mach number in the radiation direction reaches the sonic point. This is a
major limitation of the retarded time formulation. Farassat and co-workers29,30 have
recently presented a formulation that is appropriate for supersonically moving surfaces
(i.e. formulation 4) and verified by application to benchmark problems. Since, this
supersonic formulation has not yet been applied to practical problems it will not be
presented here in the interest of brevity.
The above formulation is valid when the observer is stationary and the surface is
moving at an arbitrary speed. However, for the case of an advancing blade the observer is
usually moving with the free-flow speed (e.g. rotor in a wind tunnel with a free stream not
equal to zero). The formulation can be adjusted for this case by allowing x ( t ) to move
with the free stream instead of being stationary in equation (6) for the retarded time.
It is possible to write equation (2) in a simple form valid for stationary surfaces. The
Kirchhoff formula is then

r 1 é1 Ç ¶f ù [f ]ret dS
4pf ( x , t ) = ò S r
ê f cos q -
a
ë o
ú dS +
¶n û ret ò
S r2
(7)

The retarded time for this case is t - r c. With the use of a Fourier transformation,
equation (7) can be expressed in the frequency domain (i.e. starting from Helmholtz
equation) as

r é 1 æ iw ¶fˆ ö fˆ cos q ù
4pfˆ ( x , w ) = òe cos qfˆ -
iw r / ao
ê ç- + ú dS (8)
S êë r è ao ¶n ÷ø r 2 úû
¥ 1 ¥ ˆ - iw t
fˆ =
-¥òf e iw t dt, f =
2p -¥ ò
fe dw

where f̂ is the Fourier transform of f , and w is the cyclic frequency. An equivalent to


equation (8), valid for surfaces and observers in rectilinear motion was presented by
Lyrintzis and Mankbadi34 and Pilon.35
Two-dimensional formulations can also be developed (Pilon,35 Scott et al.36). Atassi
et al.37 developed a two-dimensional frequency domain formulation that uses a
modified Green’s function in order to avoid the evaluation of normal derivatives.
Mankbadi et al.38 developed a modified Green’s function for a cylinder control surface
that was applied in jet noise predictions. Hariharan et al.39 developed a framework for
Kirchhoff’s formulations without the use of normal derivatives; the method was applied
initially for two-dimensional problems.
102 Surface integral methods in computational aeroacoustics

For completeness we should mention that for the case where the Kirchhoff control
surface S coincides with the body surface, (BIE-Boundary Integral Equations) there are
some non-uniqueness difficulties in the prediction of the radiated acoustic sound in the
exterior region whenever the frequency coincides with one of the Dirichlet
eigenfrequencies. These problems where analyzed for the stationary Kirchhoff surface
by Wu and Pierce40 and for moving Kirchhoff surfaces by Wu.41 Finally, Dowling and
Ffowcs Williams42 included the effects of infinite plane walls in the stationary
Kirchhoff formulation. However, in this paper we are reviewing the use of Kirchhoff’s
equation for extenting near-field results in the far-field (BIR Boundary Integral
Representation), so the issues mentioned in this paragraph are not relevant.

2.1.2 The extended Kirchhoff method


Equation (2) works well for aeroacoustic predictions when the control surface is placed
in a region of the flow field where the linear wave equation is valid. However, this
might not be possible for some cases. Therefore, additional nonlinearities can be added
outside the control surface.43–48 The modifications to the traditional Kirchhoff method
consist of an additional volume integral. Thus equation (2) now becomes:46 (the pressure
perturbation p ¢ is used here as the dependent variable)

r é E1 ù é p ¢E2 ù
4p p ¢ ( x , t ) = ò
S
ê ú dS +
êë r (1 - M r ) úû ret ò
S
ê 2 ú dS +
êë r (1 - Mr ) úû ret
é 1 ¶ 2 Tij ù
ò
V
ê ú dV
êë r (1 - Mr ) ¶yi ¶y j úû ret (9)
where
Tij = rui u j - s ij + ( ( p - p ) - a r ¢)d
o
2
o ij (10)

where ui is the fluid velocity, r is the density, r ¢ the density perturbation, and s ij is
the viscous stress tensor. It is easy to show that this equation reduces to the traditional
Kirchhoff integral if the control surface is placed in a fully linear region, as Tij becomes
zero. Through the use of Fourier transforms, equation (9) can also be expressed in the
frequency domain.
Isom et al.49 developed a nonlinear Kirchhoff formulation (Isom’s formulation) for
some special cases (i.e. stationary surface at the sonic cylinder of a rotor, high
frequency approximation and observer on the rotation plane). They have included in
their formulation some nonlinear effects using the transonic small disturbance equation.
The nonlinear effects are generally accounted for with a volume integral, as shown
above. However, they showed that for the above special cases the nonlinear effects can
be reduced to a surface integral.

2.2 The porous Ffowcs WilliamsÑHawkings equation


A modified integral formulation for the porous surface FW-H equation18 is needed
because the usual practice is to assume that the FW-H integration surface corresponds
to the body and is impenetrable. A convenient way to formulate this is as an extension
aeroacoustics volume 2 á number 2 á 2003 103

of Farassat’s formulation 150 which was originally developed for the rigid surface
FW-H equation. Following Di Francescantonio45 we define new variables Ui and Li as
æ rö rui
Ui = ç 1 - ÷ vi + (11)
è r oø ro

and
Li = Pij nˆ j + r ui (un - vn ) (12)

where subscript o implies ambient conditions, superscript ¢ implies disturbances


(e.g. r = r ¢ + r o ), r is the density, u is the fluid velocity, v is the velocity of the control
surface, and Pij is the compressive stress tensor with the constant pod ij subtracted. Now
by taking the time derivative of the continuity equation and subtracting the divergence
of momentum equation, followed with some rearranging, the integral form of FW-H
equation can be written as (Formulation I)
r r r r
p ¢( x , t ) = pT¢ ( x , t ) + p ¢L ( x , t ) + pQ¢ ( x , t ) (13)

where r ¶ é r oU n ù
4p pT¢ ( x , t ) =
¶t òS
ê
ë r | 1 - Mr
ú dS
| û ret
(14)

r 1 ¶ é Lr ù é Lr ù
4p p L¢ ( x , t ) =
ao ¶t ò S
ê ú dS +
ë r | 1 - Mr | û ret
òS
ê 2 ú dS
ë r | 1 - Mr | û ret
(15)

r
and pQ¢ ( x , t ) can be determined by any method currently available (e.g. references
16, 17). In equations (14) and (15) a subscript r or n indicates a dot product of the
vector with the unit vector in the radiation direction r̂ or the unit vector in the surface
normal direction n̂ i, respectively.
It should be noted that the three pressure terms have a physical meaning for rigid
r r
surfaces:
r pT¢ ( x , t ) is known as thickness noise, p L¢ ( x , t ) is called loading noise and
pQ¢ ( x , t ) is called quadrupole noise. For a porous surface the terms lose their physical
r
meaning, but the last term pQ¢ ( x , t ) still denotes the quadrupoles outside the control
(porous) surface S.
An alternative way45 is to move the time derivative inside the integral: (Formulation II)
r é ro (UÇ n + U nÇ ) ù é ro un ( r MÇ r + c ( M r - M 2 )) ù
4p pT¢ ( x , t ) = ò
S
ê 2 ú
ë r(1 - Mr ) û ret
dS +
òS
ê
ë r 2 (1 - Mr )3
ú dS
û ret
(16)

r 1 é LÇ r ù é Lr - LM ù
4p p L¢ ( x , t ) =
c ò S
ê 2ú
ë r (1 - M r ) û ret
dS +
òS
ê 2 2ú
ë r (1 - Mr ) û ret
dS

1 é Lr ( MÇ r + c( Mr - M 2 )) ù
+
c ò S
ê
ë r 2 (1 - Mr ) 3
ú dS
û ret
(17)
104 Surface integral methods in computational aeroacoustics

This is now an extension of Farassat’s formulation 1A51 (also originally developed for
the rigid surface FW-H equation) where the dot over a variable implies source-time
differentiation of that variable, LM = Li Mi , and a subscript r or n indicates a dot
product of the vector with the unit vector in the radiation direction r̂ or the unit vector
in the surface normal direction n̂ i , respectively.
Comparing the two FW-H formulations, it appears that Formulation I (equations 14, 15)
has less memory requirements, because it does not require storage of the time derivatives,
and requires less operations per integral evaluation. However, in general, integrals have
to be evaluated twice in order to find the time derivative. In the special case of a
stationary control surface, or a fixed microphone location, i.e. “flyover,” the integral can
be reused at the next time step. Since memory appears to be more important for these
type of calculations, Formulation I is a good choice for stationary surfaces. Formulation I
was used by Strawn et al.52 for rotorcraft noise predictions using a non-rotating control
surface with very good results. On the other hand, taking the time derivative inside could
prevent some instabilities. Thus Formulation II (equations 16, 17) might be more robust
for a moving control surface. Formulation II was used for rotorcraft noise prediction by
Brentner and Farassat47 with a rotating control surface with very good results. However,
a more detailed comparison of the two formulations would be very helpful.
For a stationary surface Formulation I reduces to:
r ¶ é r oU n ù
4p pT¢ ( x , t ) =
¶t ò
S êë r úû dS
ret
(18)

r 1 ¶ é Lr ù dS + é Lr ù dS
4p p ¢L ( x , t ) =
ao c ¶t ò S êë r úû
ret
òS êë r 2 úû
ret
(19)

and Formulation II becomes:


r é ro UÇ n ù
4p pT¢ ( x , t ) = ò S
ê r ú dS
ë û ret
(20)

r 1 é LÇ r ù é Lr ù
4p p L¢ ( x , t ) =
ao ò S êë r úû dS +
ret
òS êë r 2 úû dS
ret
(21)

With the use of a Fourier transformation both formulations (for a stationary surface)
can be written in the frequency domain as53

r roUˆ n
òe
iw r / ao (22)
4p pˆ T¢ ( x , w ) = - iw dS
S r
r - iw Lˆ r Lˆ r
4p pˆ ¢L ( x , w ) =
qo ò S
e iw r / ao
r
dS + òS r2
dS (23)

where pˆ ¢, Uˆ n , and L̂r are the Fourier transforms of p ¢, U n , and Lr , respectively and w is
the cyclic frequency. It should be noted that both time formulations reduce to the same
frequency formulation for a stationary control surface.
aeroacoustics volume 2 á number 2 á 2003 105

Time and frequency formulations for a uniform rectilinear motion can be found in
reference 54. Two-dimensional formulations for a solid surface FW-H equation have
already been developed in the past (see, for example, references 55, 56) and can be
readily extended to a porous surface. Finally, a supersonic formulation can also be
found in reference 33.

2.3 Comparison of Kirchhoff and FW-H methods


Both the above formulations provide a Kirchhoff-like formulation if the quadrupoles
r
outside the control surface ( pQ¢ ( x , t ) term) are ignored. The equivalence of the porous
FW-H equation and Kirchhoff formulation was proven Pilon & Lyrintzis43 and Brentner
& Farassat.44 They showed that, for a surface placed in a linear region, the porous
surface FW–H formulation is equivalent to the linear Kirchhoff formulation, plus a
volume integral of quadrupoles ( r ui u j). (Pilon and Lyrintzis43 also claim that the
control surface need not be placed in an entirely linear region. The nonlinearities can be
2
accounted for with the use of f = ao r ¢ as the dependent variable, and the volume
integral of quadrupoles, Tij .)
One difference between Kirchhoff’s and FW-H formulation is that Kirchhoff’s
method needs p ¢, ¶p / ¶n, ¶p / ¶t (3 variables) as input, whereas the porous FW-H needs
p ¢, r, rui (5 variables), or U n and Li (4 variables), or their time derivatives for
formulation II. Thus the porous FW-H method requires more memory, which can be
significant for large LES runs. The CPU time is about the same. However, the major
difference is that the porous FW-H method allows for nonlinearities on the control
surface, whereas the Kirchhoff method assumes a solution of the linear wave equation
on the surface. Thus if the solution does not satisfy the linear wave equation on the
control surface the results from the Kirchhoff method change dramatically. This leads
to a higher sensitivity for the choice of the control surface for the Kirchhoff method in
practical cases when the wave equation is not satisfied on the control surface due to
numerical errors or non-uniform velocities outside the control surface. This was shown
in reference 47 for a rotorcraft noise problem (see section 5.2). Another way to state this
difference is to state that the Kirchhoff method puts more stringent requirements to the
CFD method to reach to the linear acoustic field before dissipation and dispersion errors
due to coarsening in the far-field take over.
The volume integral of quadrupole sources that arises in the non-linear region outside
of the control surface presents a challenge. A major motivation for the use of
Kirchhoff/porous FW-H methods is the lack of volume integrations, which reduces
necessary calculations by an order of magnitude. However, the methods used in
WOPWOP+ 16,17 provide an efficient means of accounting for the quadrupoles in FW-H
calculations that may be used in both methods, because the quadrupole terms are similar.

2.4 Mean flow refraction corrections for jet noise


The Kirchhoff and the FW-H formulas presented above can efficiently and accurately
predict aerodynamically generated noise, as long as the control surface surrounds the
entire source region. In jet noise predictions, however, it is usually impossible, with
current numerical methods, to determine the entire source region. This is due to time
106 Surface integral methods in computational aeroacoustics

and memory limitations imposed by the computer architecture, as well as dispersion


and dissipation constraints. Thus, a significant nonlinear source region, as well as a
steady mean flow, will exist outside of the control surface. Even if the unsteady sound
sources outside of the control surface can be ignored, there is still a substantial steady
mean flow in the region near the jet axis, downstream of the control surface. Thus, some
means of approximating the effects of this steady shear flow are required if an acoustic
prediction is desired for observer points lying near the jet axis.
A suitable approximation to the downstream shear flow is necessary, in order to
determine the refraction effects. In the past, several researchers have used an
axisymmetric parallel shear flow model to determine sound produced by point acoustic
sources within circular jets (e.g. Amiet57). This approach was adopted by Lyrintzis and
co-workers53,58 and in order to account for refraction effects in the Kirchhoff and the
porous FW-H method. A real jet has non–zero radial velocity, but the refracting effect
of this component is minimal, and can safely be ignored. Also, the lack of azimuthal
variation in the parallel shear flow approximation has a very small effect. The value of
the axial velocity to be used in the shear flow approximation can be taken directly from
the CFD numerical simulation, at the downstream end of the control surface, as an
average of the time dependent axial velocity at each radial grid point.
The refraction problem now consists of a collection of point acoustic sources (the
integrands of equations (8) and (22) acting at radial location R, and the parallel shear flow
with U determined at each R). If the acoustic wavelength, l = 2p ao /w , is assumed to be
small compared to the shear layer thickness d , then geometric acoustics principles hold.
If the steady velocity at the downstream end of the Kirchhoff surface is denoted U s ,
the sound emission angle with respect to the jet axis J s , and the propagation angle in
the stagnant, ambient air is denoted J o , then the axial acoustic phase speeds are
preserved by the stratified flow
ao ao
P = Us + (24)
cos J o cos J s

It is assumed that the speed of sound at the source is equivalent to that in the ambient
air. This equation can be rearranged to show that there is a critical angle, J c defined by

æ 1 ö
J c = cos -1 ç ÷ (25)
è 1 + Ms ø

If the observer angle J o is greater than J c then no sound emitted at the source on the
Kirchhoff surface can reach the observer. This criterion is easily added to a stationary
surface Kirchhoff program. (Note that M s is the Mach number of the mean shear flow,
and not the Kirchhoff surface, which is assumed stationary.)
An additional correction is necessary to accurately account for the mean flow
refraction. Imposing the local “zone of silence” condition described above can allow a
surface source at a relatively large radial location to radiate sound into and through the
shear flow. This is because the local “zone of silence” decreases in size with the radial
location of the source, due to the decrease in source Mach number. The simple
aeroacoustics volume 2 á number 2 á 2003 107

correction is to set the source strength to zero if the observation point is located closer
to the jet axis than the source point on the Kirchhoff surface.
Finally, the geometric acoustics approximation is only valid for d l > 1 . It is
assumed here that the downstream end of the cylindrical Kirchhoff surface is located
far enough downstream of the jet potential core that the shear layer thickness is large
compared with the acoustic wavelength.
In reference 58 the mean flow refraction corrections were applied to the frequency
domain version of the Kirchhoff method (equation 8). In reference 53 an amplitude
correction as recommended by Amiet57 (but not included in reference 58) was added
and the methodology was applied to both Kirchhoff and FW-H methods (equations 8
and 22-23).

2.5 Open control surface


Freund et al.59 developed a way to improve the accuracy of Kirchhoff evaluations of
sound fields for an open Kirchhoff control surface. Asymptotic analysis was used to
provide correction terms which partially account for the missing portion of the integral
surface. It was shown that the major contribution comes from a point on the surface that
intersects the line between observer and source. A correction term was estimated to
account for the missing parts of the Kirchhoff surface. The study is restricted to the case
where the mean flow is parallel to the available surface, as happens for example, for
jet noise problems when the downstream surface vertical to the jet axis is missing.
The corrections are limited to observers away from the jet axis. More details can be
found in the original reference. It appears that this study can be extended to the porous
FW-H equation, as well.

3 ALGORITHMIC ISSUES
Some algorithmic issues are discussed below. Additional information for numerical
algorithms for acoustic integrals, in general, is given by Brentner.60

3.1 Choice of control surface


The Kirchhoff scheme requires stored data for pressure and pressure derivatives on a
surface. Since Kirchhoff’s method assumes that the linear wave equation is valid
outside the closed control surface S, S must be chosen large enough to include the
region of all nonlinear behavior. However, the accuracy of the numerical solution is
limited to the region immediately surrounding the moving blade because of the increase
of mesh spacing in CFD codes. Thus a judicious choice of S is required for the
effectiveness of the Kirchhoff method. For example, in the case of airfoil/rotor noise the
control surface is typically located a couple of chordlengths away from the airfoil/rotor
surface.
For a porous FW-H formulation no normal derivatives are required and (because
nonlinearities are allowed on the control surface) the results are less sensitive to the
choice of the control surface,47 as will be shown in section 5.2. Thus the CFD
requirements for the FW-H are less strigent, making the method more attractive. Singer
et al.61 used a FW-H method for the analysis of slat trailing-edge flow. The interesting
108 Surface integral methods in computational aeroacoustics

thing about this application is that part of the control surface is solid and another part
is porous.

3.2 Quadrature
For sufficient accuracy in the far-field calculations, high order quadrature should
be used to solve the surface integrals in equation (2). The predicted surface quantities
( p ¢, ¶p ¶n , ¶p ¶t , r, rui ) should also be very accurate. This can be achieved through
the use of a very fine mesh in the CFD calculations. However, memory and time
constraints often make this impractical. Meadows and Atkins62 have shown that it is
possible to obtain highly accurate Kirchhoff predictions from relatively coarse–grid
CFD solutions. Through an interpolation process, more spatial points are added to the
Kirchhoff quadrature calculations without additional effort in the CFD process. This has
the effect of refining the CFD mesh with almost no additional cost. They refer to this
process as “enrichment”. High order quadrature, temporal interpolation, and enrichment
are important for accurate far-field noise predictions for both the Kirchhoff and the
FW-H equation methods, especially if the CFD grid resolution is somewhat coarse.

3.3 Retarded or forward time


The retarded time equation (5) has a unique solution when the surface moves
subsonically. A Newton-Raphson (or divide and conquer) method can be used to solve
this nonlinear equation. This method has been the basis of several Kirchhoff codes
(e.g. Lyrintzis & Mankbadi34, Strawn et al.63, Lyrintzis et al.64 Polacsec & Prier65). The
algorithm can be easily parallelized (e.g. Wissink et al.66, Strawn et al.67) by partitioning
the control surface and distributing to different processors. Since the only
communication is the final global summation the parallel efficiency of the code is very
high. Lockard54 discussed parallelization of FW-H codes. Long and Brentner68
proposed a master-slave approach for load balancing.
However, it is difficult to write a versatile code for various mesh topologies used by
current CFD codes, including unstructured grids, based on this approach. In addition,
when these codes are extended to supersonically moving surfaces, the retarded time
equation will have multiple roots that will be difficult to evaluate. Also, the codes
sometimes require significant memory. Finally62, the variation of the source strength
over a surface element in the retarded time can be very high at certain observer
locations ( rˆ × nˆ ® 0 ) and near sonic velocities ( Mr ® 1) requiring a large number of
points per wavelength.
In order to overcome the limitations stated above, another approach can be
developed which accumulates signals (with time moving forward) from each surface
element to an observer, thus it avoids the retarded time calculation. Computer memory
requirements are reduced dramatically and the algorithm is inherently parallel. In this
approach, the final overall observer acoustic signal is found from the summation of the
acoustic signal radiated from each source element of control surface during the same
source time. The observer time is a straight-forward calculation using equation (6).
For each surface element time is moved forward from the source (emission) to the
observer time. Since a different surface element will result in a different observer time,
aeroacoustics volume 2 á number 2 á 2003 109

interpolation techniques are required when the integration is performed to obtain the
overall acoustic signal at the observer position. Both linear interpolation and spline
subroutines can be used. For high frequencies a digital filter may be used to increase
accuracy (e.g. Glegg69). This method has been used by several investigators,
e.g. Ozyoruk and Long,70–72 Lyrintzis and Xue,73 and Rahier and Prier,74 Algermissen
and Wagner,75 Delriex et al.,76 and Kim et al.77 Finally, a marching-cubes algorithm78
can be used to provide an efficient algorithm that is easy to parallelize for the evaluation
of the propagation from an emission surface.

3.4 Rotating or nonrotating control surface


For rotor applications both a rotating and a nonrotating formulation can be used.
A non-rotating formulation uses a nonrotating control surface that encloses the entire rotor
(e.g. Forsyth and Korkan,79 Strawn and Biswas,80 Baeder et al.81). A rotating Kirchhoff
formulation allows the control surface to rotate with the blade aligning with the CFD lines,
e.g. Xue and Lyrintzis,82 Lyrintzis et al.,64 Polacsec and Prier.65 No transformation of data
is needed since the CFD input is also rotating. A comparison of the rotating and the
nonrotating Kirchhoff methods showed that both methods are very accurate and efficient
(Strawn et al.63). For the porous FWH method there are fewer applications. A rotating
method was used in references 45, 47, 76 and 83 and a nonrotating method in reference 52.
It should be noted that the nonrotating formulation requires reliable data out to a
nonrotating cylinder (i.e. the control surface) surface that is usually farther out than a
rotating surface. Therefore, more accuracy of the CFD results is needed. Thus the
nonrotating method has been used in conjunction with Euler/Navier Stokes codes (e.g.,
TURNS code,84,85 OVERFLOW code86) whereas the rotating Kirchhoff method has
been used with full potential codes (e.g. FPR code87,88), as well. However, a drawback
of the rotating method is that the rotating speed of the tip of the rotating surface needs
to remain subsonic, to use the subsonic formulas shown in section 2, because of
the singularities appearing at some terms. For high tip speeds (e.g. Mtip = 0.92 ) the
supersonic formulation of Farassat et al.32,33 can be employed. For forward time
algorithms the singularities disappear and a simple method for supersonic rotation
speeds has been developed by Delrieux et al.76

4 VALIDATION RESULTS
Both Kirchhoff and FWH formulations have been validated using model problems.
The first thing to do is, of course, check that the signal becomes zero inside the control
surface. The number of points per period and the number of points per wave length
should also be studied.34,53
A stationary or translating point source have been used by Lyrintzis et al.,34,53
Myers & Hausmann,31 and Lockard54 and a rotating point source by Lyrintzis et al.64
and Berezin et al.89 Exponential source distributions have been used by Pilon and
Lyrintzis.35,43,44,46 Hu et al.90 used a line monopole source and a Gaussian pressure and
vorticity pulse (category 3 benchmark problem91) to verify their two-dimensional
FW-H formulation. Farassat and Farris 33 used dipole distributions on a flat surface and
a sphere to validate the supersonic formulation (i.e. formulation 4). Singer et al.92 used
110 Surface integral methods in computational aeroacoustics

a line vortex around an edge. Meadows and Atkins62 used an oscillating sphere and
studied the effects of quadrature (see section 3). Ozyoruk and Long68 have used the
scattering problem of sound by a sphere (Figure 2). The spherical sound waves are
generated by a partially distributed Gaussian mass source. The results from an exact
solution and a direct Euler solver are also shown. Note that near 180° the Kirchhoff
results are better than the direct calculation, because of numerical dissipation as the
waves travel longer distances to arrive at the observer locations.

5 AEROACOUSTIC APPLICATIONS
Kirchhoff’s formula has been extensively used in light diffraction and other
electromagnetic problems, aerodynamic problems, i.e. boundary-elements (e.g. Morino
et al.28), as well as in problems of wave propagation in acoustics (e.g. Pierce93).
Kirchhoff’s integral formulation has been used extensively for the prediction of
acoustic radiation in terms of quantities on boundary surfaces (the Kirchhoff control
surface coincides with the body). Kirchhoff’s method has also been used for the
computation of acoustic scattering from rigid bodies using a boundary element
technique with the Galerkin method. The solid surface FW-H equation with its various
forms21 has been used in several problems including propeller and helicopter noise.
Here we will concentrate on the use of “Kirchhoff,” and “porous” FW-H equation
methods, i.e. using a nonlinear CFD solver for the evaluation of acoustic sources in
the near-field and a Kirchhoff/porous FW-H formulation for the acoustic propagation.
We will review some “real-life” aeroacoustic applications of both methods con-
centrating in recent advances.

5.1 Propeller noise


Hawkings18 suggested a stationary-surface Kirchhoff formula to predict the noise
from high-speed propellers and helicopter rotors. Forsyth and Korkan79 calculated

DIRECTIVITY AT 4.5 SPHERE RADII


0.50
Exact solution
Normalized RMS Pressure

Direct CFD data


0.40 Kirchhoff prediction using CFD data

0.30

0.20

0.10
0 50 100 150
Angle From Negative x-Axis, Degrees
Figure 2. Sound scattering by a sphere. Comparison with exact solution (from
reference 70).
aeroacoustics volume 2 á number 2 á 2003 111

high-speed propeller noise using the Kirchhoff formulation of Hawkings.18 Jaeger and
Korkan94 used a special case of the Farassat and Myers25 formulation for a uniformly
moving surface to extend the calculation to advancing propellers. In the above
applications, the control surface S was chosen to be a cylinder enclosing the rotor.

5.2 Helicopter impulsive noise


Kirchhoff’s method has been widely applied in the prediction of helicopter impulsive
noise.95 The Kirchhoff method for a uniformly moving surface was initially used in
two-dimensional transonic Blade-Vortex Interactions (BVI) to extend the numerically
calculated nonlinear aerodynamic BVI results to the linear acoustic far-field.96–98
Actually, the first application of Hawkings18 Kirchhoff Method was given by George and
Lyrintzis,19 where the terminology “Kirchhoff Method” was introduced. The Kirchhoff
method was used to test ideas for BVI noise reduction (Xue and Lyrintzis99). The method
was also extended to study noise due to other unsteady transonic flow phenomena
(i.e. oscillating flaps, thickening-thinning airfoil) by Lyrintzis et al.100 Later, the method
was used for the two-dimensional BVI problem by Lin and co-workers.101,102
Kirchhoff’s method has also been applied to three-dimensional High-Speed
Impulsive (HSI) noise. Baeder et al.81 and Strawn & Biswas80 used a nonrotating
control Kirchhoff surface that encloses the entire rotor. The Transonic Unsteady Rotor
Navier Stokes (TURNS) code84,85 was used for the near-field CFD calculations.
An unstructured grid was used by Strawn et al.103 and an overset grid code (OVERFLOW)86
by Ahmad et al.104 Kirchhoff’s method predicted the HSI hover noise very well using a
small fraction of CPU time of the straight CFD calculation.
Another Kirchhoff method used in helicopter noise is the rotating Kirchhoff method
(i.e. the surface rotates with the blade). The method was used for three-dimensional
transonic BVI’s for a hovering rotor by Xue and Lyrintzis.82 The near-field was
calculated using the Full Potential Rotor (FPR) code.87,88 The rotating Kirchhoff
formulation allows the Kirchhoff control surface to rotate with the blade; thus a smaller
cylinder surface around the blade can be used. No transformation of data is needed
because the CFD input is also rotating. Since more detailed information is utilized for
the accurate prediction of the far-field noise this method is more efficient. Finally, the
method was extended for an advancing rotor and was applied to HSI noise105 and BVI
noise.106,107 Berezin et al.89 showed that sometimes special care is needed for choosing
the CFD grids, because the highly stretched grids used for aerodynamic applications
may not provide accurate information on the control Kirchhoff surface.
A comparison63 of the rotating and the nonrotating Kirchhoff methods showed that
both methods are very accurate and efficient. Figure 3 shows a comparison for an
advancing HSI noise case (1/7 scale AH-1 helicopter, hover tip Mach number
M H = 0.665, advance ratio m = 0.258, which corresponds to an advancing tip Mach
number of M at = 0.837). TURNS84,85 is used for the CFD calculations. We see that
both methods compare very well with the experiments.108 Kirchhoff’s method has
become a standard tool for rotorcraft acoustic predictions. The method is currently
implemented in the TRAC (TiltRotor Aeroacoustic Codes) system developed by NASA
Langley (RKIR code, Lyrintzis et al.,64 Berezin et al.89) and is employed at NASA Ames
112 Surface integral methods in computational aeroacoustics

AFDD (Strawn et al.63). In Europe, additional versions of rotating and nonrotating


Kirchhoff codes have also been developed.65,74,75,76,109,110
Kirchhoff’s method results have also been compared with the traditional acoustic
analogy (solid surface FW-H equation). A comparison with the acoustic analogy code
WOPWOP 15 (WOPWOP uses the solid surface FW-H equation without accounting for
quadrupoles) has shown that Kirchhoff method is superior when quadrupole sources are
present (Lyrintzis et al.111) for advancing HSI cases. Baeder et al.81 also compared the
results with a linear (i.e. monopole plus dipole sources on the rotating blade) solid
surface FW-H equation method for hover HSI. The FW-H results were inaccurate for
tip Mach numbers higher than 0.7, because of the omission of quadrupole sources.
However, a further comparison of the rotating Kirchhoff method to WOPWOP+ 16,17
(WOPWOP+ is a solid surface FW-H equation method accounting also for quadrupoles
with a volume integral) has shown that the two methods give about the same results
(Brentner et al.112), but Kirchhoff method uses only surface integrals and avoids the
quadrupole volume integration. It should be noted that the robustness of the Kirchhoff
method improves with the use of a less stretched grid (Berezin et al.89) or an Euler code,
e.g. TURNS (Baeder et al.81).

20
Experiment
Rotating Kirchhoff
0
Nonrotating Kirchhoff
Pressure (Pa)

2 20

2 40
r/ R5 6.88 4
2 60 (2) (3)

2 80
240 250 260 270 280 210 220 230 240 250

20
r/ R5 3.44
Pressure (Pa)

2
0 1 3

2 20

2 40 (1) (4)
30° 2 30°

2 60
260 270 280 290 300 210 220 230 240 250
Blade azimuthal angle (deg) w Blade azimuthal angle (deg)

Figure 3. Comparison of acoustic pressures with experimental data102 at four


different microphone locations for an AH-1 blade with M at = 0.837. All
microphones are in the plane of the rotor (from reference 63).
aeroacoustics volume 2 á number 2 á 2003 113

Isom et al.,49 and Purcell113,114 used a modified Kirchhoff method which also
included some nonlinear effects for a stationary surface, to calculate hover HSI noise.
Results (not shown here) show good agreement with experimental data.
A porous FW-H method based on Kirchhoff subroutines was also developed by
Brentner & Farassat47 (FWH/RKIR code), Morgans et al.,83 Strawn et al.,52 and Delriex
et al.76 These codes do not include quadrupoles outside the control surface, because it
was found to be of minor importance unless the Mach number is really high.115 Thus
the porous FW-H equation is also based on surface integrals. The porous FW-H
formalism is more robust than the traditional Kirchhoff method with regards to the
choice of the control surface, as shown in Figures 4 and 5 for a hover HSI noise case
(1/4 model UH-1H model helicopter, hovering at M H = 0.88 , experiments from
Purcell113). FPR87,88 was used for the CFD calculations.

5.3 Airfoils
Atassi and his co-workers37,116–118 have used Kirchhoff’s method for the evaluation of
acoustic radiation from airfoils in nonuniform subsonic flows. They employed rapid
distortion theory to calculate the near-field CFD. A sample comparison for the far-field
directivity of the acoustic pressure using the Kirchhoff method and the direct calculation
method (i.e. rapid distortion theory119–121 is given in Figure 6 (from reference 37) for
a 3% thick Joukowski airfoil in a transverse gust at k1 = (w c / 2 V¥ ) = 1 and M = 0.1.
The semi-analytical results for a flat plate encountering the same gust are also shown
in Figure 6 and are very close to the results from Kirchhoff’s method. The figure
indicates that the direct calculation method is not accurate in the far-field, as the direct

2 1000 k=2
k=7
k=12
2 2000
k=18
p¢, Pa

k=21
2 3000
data

2 4000

2 5000
0.0 0.5 1.0 1.5 2.0
time, msec
Figure 4. Comparison of Kirchhoff acoustic pressures with experimental data108
for an observer in the plane of the rotor at 3;4R from a UH-1H model
rotor hovering at M H = 0.88 (from reference 47).
114 Surface integral methods in computational aeroacoustics

100

0
k=2
k=7
2 100
k=12
p¢, Pa

k=18
2 200 k=21
data

2 300

2 400
0.0 0.5 1.0 1.5 2.0
time, msec
Figure 5. Comparison of porous FW-H acoustic pressures with experimental
data108 for an observer in the plane of the rotor at 3, 4R from a UH-1H
model rotor hovering at M H = 0.88 (from reference 47).

simulation results are very different from the semi-analytical and the Kirchhoff
results. This is due to discretization errors. However, this CFD code is accurate in the
near-field and the Kirchhoff method should be used instead in the far-field, as indicated
in Figure 6.
Singer et al.92,61 used a FW-H method for the evaluation of acoustic scattering from
a trailing edge and slat trailing edge. The interesting thing about the slat trailing edge
application is that part of the control surface is solid and another part is porous.

5.4 Fan noise


Kirchhoff’s method can also be applied to ducted fan noise. Very good results were
shown by Ozyoruk and Long70–72 for a control surface in rectilinear motion. A forward
time parallel algorithm was used. A porous FWH method was used by Zhang122 with
very good results.

5.5 Jet noise


Kirchhoff’s method has also been applied in the estimation of jet noise. Soh,123 Mitchel
et al.124 Zhao et al.,125 and Billson et al.126 used the stationary Kirchhoff method
(equation 7) and Lyrintzis & Mankbadi,34 Chyczewski & Long,127 Morris et al.,128
Gamet and Estivalezes,129 Choi et al.,130 and Kandula and Caimi131 used the uniformly
moving formula. It should be noted that most of the above references use an LES code for
the CFD data. However, a RANS code can also be used, as shown in reference 131, where
OVERFLOW86 was used. Lyrintzis & Mankbadi34 also compared time and frequency
domain formulations. Mankbadi et al.38 applied a modified Green’s function to avoid
aeroacoustics volume 2 á number 2 á 2003 115

0.20

0.15

0.10

0.05

0
2 0.15 2 0.10 2 0.05 0 0.05 0.10 0.15 0.20
Figure 6. Comparison between far-field directivity of acoustic pressure values
using the Kirchhoff method (- -) and the direct calculation method (-•-)
for a 3% thick Joukowski airfoil in a transverse gust at k1 = 1.0, M = 0.1.
The semi analytical results (–) for a flat plate encountering the same gust
are also shown (from reference 37).

the evaluation of normal derivatives. Balakumar132 and Yen133 used parabolized stability
equations for the jet simulation and a cylindrical (i.e. two-dimensional) Kirchhoff
formulation for the noise evaluation. Shih et al.134 compared several Kirchhoff
formulations with the acoustic analogy, extending the LES calculations and using a zonal
LES + LEE method. The results showed that the Kirchhoff method is much more accurate
than the acoustic analogy (for the compact source approximation used) and much cheaper
than extending the LES or performing a zonal LES + LEE.
The uses of FW-H method in jet noise have been sparse. Morris et al.135,136 and Uzun
et al.137 used the method with good results and Hu et al.,90,138 used a two-dimensional
formulation of the porous FW-H equation to evaluate noise radiation from a plane
jet. Rahier et al.139 compared the methods for numerical acoustic predictions for hot
jets. Both the Kirchhoff method based on pressure disturbance and the FW-H equation
gave good results, whereas the Kirchhoff method based on density gave erroneous
results.
Most of the above approaches have used an open control surface (i.e. without
the downstream end) in order to avoid placing the surface in a nonlinear region. Freund
et al.59 showed a means of correcting the results to account for an open control surface,
for cases that the observer is close to the jet axis. Pilon and Lyrintzis43,44,47 developed
a method to account for quadrupole sources outside the control surface. This
approximation is based on the assumption that all wave modes approximately decay
in an exponential fashion. The volume integral is reduced to a surface integral for a
far-field low frequency approximation and a Taylor series expansion for axisymmetric
jets. However, a simpler method (suggested in reference 53) is to just use an existing
empirical code (e.g. MGB140) to evaluate the noise using as inflow the CFD solution on
the right side of the control surface. Thus MGB can provide an estimate of the error of
ignoring any sources outside the control surface of the Kirchhoff/porous FW-H method.
116 Surface integral methods in computational aeroacoustics

An approximate way to account for refraction effects was developed by Lyrintzis


and co-workers,53,58 as explained above in section 2.5. A typical result shown
here (Figure 7) shows the effects of refraction corrections for a supersonic Mach
number case (excited, Mach 2.1, unheated ( To = 294 K ), round jet of Reynolds Number
Re = 70000; the jet exit variables were perturbed at a single axisymmetric mode
at a Strouhal number of St = 0.20, the amplitude of the perturbation was 2% of the
mean). Further development of refraction corrections based on Lilley’s equation,
(e.g. reference 141) is possible.
Finally, it should be noted that for some complicated noise problems (as, for example,
in jet noise) several computational domains might be needed: a complicated near-field
(e.g. using Large Eddy Simulations-LES), a simplified mid-field with some nonlinear
effects, and a linear Kirchhoff’s method for the far-field. Kirchhoff’s formulation can
be the simplest region of a general zonal methodology. This idea has been proposed by
Lyrintzis,19 but it has not yet been implemented.

5.6 Other applications


Other applications of these surface integral methods have also been attempted. For
example, in reference 142 a Kirchhoff method was applied to a cavity problem, with
application to vehicle noise predictions. Also the method is now part of the Fluent code
and has been applied to a wide range of problems77 and is currently implemented143 in
the Star-CD code.

50
No Corrections
40

30

20

10
R/Rj

2 10

2 20

2 30

2 40
Refraction Corrections
2 50
0 10 20 30 40 50 60 70 80 90 100
x/Rj

Figure 7. Instantaneous contours of ao2 r ¢ po . R > 0 : No refraction corrections.


R < 0 : Refraction corrections imposed (from reference 58).
aeroacoustics volume 2 á number 2 á 2003 117

Finally, it should be noted that FW-H can not be used with incompressible CFD data.
For low speeds (i.e. incompressible flow) and a stationary, impermeable surface the
FW-H equation reduces to Curle’s integral (reference 144). Wang et al.145 used Curle’s
integral for the evaluation of trailing edge noise with an incompressible Navier Stokes
solver. (They concluded that the volume integral of the equation, i.e. quadrupoles
outside the airfoil surface, is not important for low speeds.)

6 CONCLUDING REMARKS
Kirchhoff’s and porous FW-H methods consist of the calculation of the nonlinear near-
and mid-field numerically with the far-field solutions found from a Kirchhoff/porous
FW-H formulation evaluated on a control surface S surrounding the nonlinear-field.
The surface S is assumed to include all the nonlinear flow effects and noise sources. The
separation of the problem into linear and nonlinear regions allows the use of the most
appropriate numerical methodology for each. The advantage of these methods is that
the surface integrals and the first derivatives needed can be evaluated more easily than
the volume integrals and the second derivatives needed for the evaluation of the
quadrupole terms when the traditional acoustic analogy is used.
The porous FW-H equation method is a newer idea and there fewer applications in
the literature. The method is equivalent to Kirchhoff’s method and has the same
advantanges. In comparison, it is more robust with the choice of control surface and
does not require normal derivatives. Since the method also requires a surface integral,
it is very easy to modify existing Kirchhoff/solid surface FW-H codes. The method
requires larger memory, because more quantities on the control surface are needed.
However, we believe that the robustness is more important and thus the porous FW-H
is the method we recommend.
The use of both methods has increased substantially over the last 10 years, because
of the development of reliable CFD methods that can be used for the evaluation of the
near-field. The methods can be used to study various acoustic problems, such as
propeller noise, high-speed compressibility noise, blade-vortex interactions, jet noise,
ducted fan noise, etc. Some results indicative of the uses of both methods are shown
here, but the reader is referred to the original references for further details. The methods
are becoming more popular and have been coupled with production codes, such as
OVERFLOW, Fluent and Star-CD. We believe that, a simple set of portable
Kirchhoff/FW-H subroutines can be developed to calculate the far-field noise from
inputs supplied by any aerodynamic near/mid-field code.

ACKNOWLEDGEMENTS
The author was supported by the Indiana 21st Century Research and Technology Fund,
and the Aeroacoustics Research Consortium (AARC), a government and industry
consortium mananged by the Ohio Aerospace Institute (OAI).

REFERENCES
1. Lighthill, M. J., “A General Introduction to Aeroacoustics and Atmospheric Sound,”
ICASE Report 92-52, NASA Langley Research Center, Hampton, VA 1992.
118 Surface integral methods in computational aeroacoustics

2. Hardin, J., and Hussaini, M. Y., Computational Aeroacoustics, Springer-Verlag,


New York, 1993.
3. Tam, C. K. W., “Computational Aeroacoustics: Issues and Methods,” AIAA
Journal, Vol. 33, No. 10, Oct. 1995, pp. 1788–1796.
4. Long, L. N., Morris, P. J., Ahuja, V., Hall, C., and Liu, J., “Several Aerospace
Applications of Computational Aeroacoustics,” Proceedings of the ASME Noise
Control and Acoustics Division, NCA-Vol. 25, the 1998 ASME International
Mechanical Engineering Congress and Exposition, Nov. 1998, Anaheim, CA,
pp. 13–20.
5. Mankbadi, R. R., Hayder, M. E., Povinelli, L. A., “The Structure of Supersonic Jet
Flow and Its Radiated Sound,” AIAA Journal, 32, 1994, pp. 897–906.
6. Stoker, R. W., and Smith, M. J., “An Evaluation of Finite Volume Direct
Simulation and Perturbation Methods in CAA Applications,” AIAA 93-0152,
Aerospace Sciences Meeting, Reno, NV, Jan. 1993.
7. Lim, T. B., Sankar, L. N., Hariharan, N., and Reddy, N. N., “A Technique for the
Prediction of Propeller Induced Acoustic Loads on Aircraft Structures,” AIAA-
93-4340, 15th AIAA Aeroacoustics Conference, Long Beach CA.
8. Viswanathan, K., and Sankar, L. N., “Toward the Direct Calculation of Noise:
Fluid/Acoustic Coupled Simulation,” AIAA Journal, Vol. 33, No. 12, pp. 2271–2279.
9. Shih, S. H., Hixon, D. R., and Mankbadi, R. R., “A Zonal Approach for the
Prediction of Jet Noise,” CEAS/AIAA paper 95-144 presented at the first joint
CEAS/AIAA Aeroacoustics Conference (16th AIAA Aeroacoustics Conference),
Munich Germany, June 1995.
10. Hardin, J. C., and Pope D.S., “Sound Generated by Flow over a Two-Dimensional
Cavity,” AIAA Journal Vol. 33, No. 3, March 1995, pp. 407–412.
11. Freund J.B., “A Simple Method for Computing Far-Field Sound In Aeroacoustic
Calculations, Journal of Computational Physics, Vol. 157, 2000, pp. 796–800.
12. Lighthill, M. J., “On the Sound Generated Aerodynamically, Part I, General
Theory,” Proc. R. Soc. London A 211, pp. 567–587.
13. Lilley, G. M., “On the Noise from Jets: Noise Mechanisms,” AGARD-CP-131,
March 1974, 13.1–13.12.
14. Ffowcs Williams, J. E., and Hawkings, D. L., “Sound Generation by Turbulence
and Surfaces in Arbitrary Motion,” Philosophical Transactions of the Royal
Society of London, Vol. 264A, May 1969, pp. 321–342.
15. Brentner, K. S., “Prediction of Helicopter Rotor Discrete Frequency Noise,”
NASA Technical Memorandum 87721, Oct. 1986.
16. Brentner, K. S., and Holland, P. C., “An Efficient and Robust Method for
Computing Quadrupole Noise,” Journal of the American Helicopter Society,
Vol. 42, No. 2. Apr. 1997, pp. 172–181.
17. Brentner, K. S., “An Efficient and Robust Method for Predicting Rotor High-Speed
Impulsive Noise,” Journal of Sound and Vibration, Vol. 203 (1), 1997, pp. 87–100.
aeroacoustics volume 2 á number 2 á 2003 119

18. Hawkings, D. L., “Noise Generation by Transonic Open Rotors,” Westland


Research Paper 599, 1979.
19. George, A. R., and Lyrintzis, A. S., “Mid-Field and Far-Field Calculations of
Transonic Blade-Vortex Interactions,” AIAA paper 86-1854, AIAA 10th
Aeroacoustics Conference, Seattle, WA, July 9-11, 1986.
20. Lyrintzis, A. S., “Review: The Use of Kirchhoff Method in Aeroacoustics,” ASME
Journal of Fluids Engineering, Vol. 116, No. 4, Dec. 1994, pp. 665–676.
21. Crighton, D. G., Dowling, A. P., Ffowcs Williams, J. E., Heckl, M., and
Leppington, F. G., Modern Methods in Analytical Acoustics: Lecture Notes,
Springer–Verlag, London, 1992.
22. Farassat, F., “Acoustic Radiation From Rotating Blades - The Kirchhoff Method
in Aeroacoustics, Journal of Sound and Vibration, Vol. 239, No. 4, January, 2001,
pp. 785–800.
23. Kirchhoff, G. R., “Zur Theorie der Lichtstrahlen,” Annalen der Physik und
Chemie, Vol. 18, 1883, pp. 663–695.
24. Morgans, R. P., “The Kirchhoff Formula Extended to a Moving Surface,”
Philosophical Magazine, Vol. 9, s.7, No. 55, 1930, pp. 141–161.
25. Farassat, F., and Myers, M. K., “Extension of Kirchhoff’s Formula to Radiation
from Moving Surfaces,” Journal of Sound and Vibration, Vol. 123, No. 3, 1988,
pp. 451–461.
26. Morino, L., “A General Theory of Unsteady Compressible Potential
Aerodynamics,” NASA Contractor Report CR-2464, December 1974.
27. Morino, L., Bharadvaj, B. K., Freedman, M. I., and Tseng, K., “BEM For Wave
Equation With Boundary in Arbitrary Motion And Applications to Compressible
Potential Aerodynamics of Aeroplanes and Helicopters” in Computational
Mechanics, Vol. 4, 1989, pp. 231–243.
28. Gennaretti, M., Luceri, L., and Morino, L., “A Unified Boundary Integral
Methodology for Aerodynamics and Aeroacoustics of Rotors,” Journal of Sound
and Vibration, Vol. 200, 1997, pp. 467–489.
29. Morino, L., Bernardini, G., and Gennaretti, M., “A Velocity-Potential-Based
Boundary-Element Method for the Aeroacoustics Analysis of Rotors and
Propellers in Arbitrary motion,” AIAA paper 2002-2539, presented at the 8th
AIAA/CEAS Aeroacoustics Conference, Breckenridge CO, June 2002.
30. Morino, L., “Is there a Difference Between Aeroacoustics and Aerodynamics? An
Aeroelastician’s viewpoint,” AIAA Journal, Vol. 41, No. 7, July 2003, pp.
1209–1223.
31. Myers, M. K., and Hausmann, J. S., “On the Application of the Kirchhoff Formula
for Moving Surfaces,” Journal of Sound and Vibration, Vol. 139, 1990, pp. 174–178.
32. Farassat, F., Brentner, K. S., and Dunn, M. H., “A Study of Supersonic Surface
Sources - The Ffowcs Williams Hawkings Equation and the Kirchhoff Formula,”
AIAA paper 98-2375, Proceedings of the 4th AIAA/CEAS Aeroacoustics
Conference, Toulouse, France, June 1998.
120 Surface integral methods in computational aeroacoustics

33. Farassat, F., and Farris, M., “Verification and Analysis of Formulation 4 of
Langley for the Study of Noise From High Speed Surfaces,” AIIA paper 99-1881
5th AIAA/CEAS Aeroacoustics Conference, Bellevue, WA, May, 1999.
34. Lyrintzis, A. S., and Mankbadi, R. R., “On the Prediction of the Far-Field Jet
Noise Using the Kirchhoff Method,” AIAA Paper 95-0508, presented at the 33rd
AIAA Aerospace Sciences Conference, Reno, NV, Jan. 1995; also AIAA Journal,
Vol. 34, Feb. 1996, pp. 413–416.
35. Pilon, A.R., Development of Improved Surface Integral Methods for Jet
Aeroacoustic Predictions, PhD Thesis, Dept. of Aerospace Engineering and
Mechanics, University of Minnesota, Minneapolis, MN, January 1997.
36. Scott, J. N., Pilon, A. R., Lyrintzis, A. S., and Rozmajzl, T., “A Numerical
Investigation of Noise from a Rectangular Jet,” AIAA paper No. 97-0485,
presented at the 35th Aerospace Science Meeting, Reno, NV, Jan. 1997.
37. Atassi, H. M., Subramaniam, S., and Scott, J. R., “Acoustic Radiation from
Lifting Airfoils in Compressible Subsonic Flow,” NASA Technical Memorandum
103650; also AIAA paper 90-3911, Oct. 1990.
38. Mankbadi, R. R., Shih, S. H., Hixon, R., Stuart, J. T., and Povinelli, L. A.,
“Extension of Near Field to Far Field Noise Prediction,” AIAA paper 96-2651, 32
Joint Propulsion Conference Lake Bueva Vista, FL, July 1996.
39. Hariharan, S. I., Scott, J. R., and Kreider, K. L., “Potential Theoretic Methods for
Far-Field Sound Radiation Calculations,” Journal of Computational Physics Vol.
164, 2000, pp. 143–164.
40. Wu, S. F., and Pierce, A. D., “Nonuniqueness of Solutions to Variationally
Formulated Acoustic Radiation Problems,” ASME Journal of Vibration and
Acoustics, Vol. 112, 1990, pp. 263–267.
41. Wu, S. F., “Nonuniqueness of Solutions to Extended Kirchhoff Integral
Formulations,” Journal of the Acoustical Society of America, Vol. 93, No. 2,
February, 1993, pp. 683–695.
42. Dowling, A. P., and Ffowcs Williams, Sound and Sources of Sound, Wiley &
Sons, New York, 1982.
43. Pilon, A., and Lyrintzis, A. S., “On the Development of a Modified Kirchhoff
Method for Supersonic Jet Aeroacoustics,” AIAA paper No. 96-1709, presented at
the 2nd AIAA/CEAS Aeroacoustics Meeting, (17th AIAA Aeroacoustics
Meeting) State College, PA, May 1996.
44. Pilon, A. R., and Lyrintzis, A. S., “Integral Methods for Computational
Aeroacoustics,” AIAA paper No. 97-0020, presented at the 35th Aerospace
Science Meeting, Reno, NV, Jan. 1997.
45. Di Francescantonio, P., “A New Boundary Integral Formulation for the Prediction
of Sound Radiation,” Journal of Sound and Vibration Vol. 202, No. 4, 1997, pp.
491–509.
46. Pilon, A. R., and Lyrintzis, A. S., “Development of an Improved Kirchhoff Method
for Jet Aeroacoustics,” AIAA Journal, Vol. 36, No. 5, May 1998, pp. 783–790.
aeroacoustics volume 2 á number 2 á 2003 121

47. Brentner, K. S., and Farassat, F., “An Analytical Comparison of the Acoustic
Analogy and Kirchhoff Formulations for Moving Surfaces,” AIAA Journal, Vol. 36,
No. 8, Aug. 1998, pp. 1379–1386.
48. Prier, J., and Rahier, G., “Aeroacoustic Integral Methods and Efficient Numerical
Implementation,” Aerospace Science and Technology Vol. 5, No. 7, pp. 457–468,
Oct. 2001.
49. Isom, M. P., Purcell, T. W., and Strawn, R. C., “Geometrical Acoustics and
Transonic Sound,” AIAA paper 87-2748, AIAA 11th Aeroacoustics Conference,
Sunnyvale, CA. 1987.
50. Farassat F., “Linear Acoustic Formulas for Calculation of Rotating Blade Noise,”
AIAA Journal, Vol. 19, No. 9, Sept. 1981, pp. 1122–1130.
51. Farassat, F., and Succi, G. P., “The Prediction of Helicopter Rotor Discrete
Frequency Noise,” Vertica, Vol. 7, No. 4, 1983, pp. 309–320.
52. Strawn, R. C., Ahmad, J., and Duque, E. P. N., “Rotorcraft Aeroacoustics
Calculations with Overset-Grid CFD Methods,” Journal of American Helicopter
Society, Vol. 44, No. 2, April 1999, pp. 132–140.
53. Lyrintzis, A. S. and Uzun, A., “Integral Techniques for Jet Aeroacoustics
Calculations,” AIAA paper 2001-2253 presented at the 7th AIAA/CEAS
Aeroacoustics Conference, Maastricht, Netherlands, May 2001.
54. Lockard, D., “A Comparison of Ffowcs Williams-Hawkings Solvers for Airframe
Noise Applications, AIAA paper No. 2002-2580 presented at the 8th AIAA/CEAS
Aeroacoustics Conference, Breckenridge, CO, June 2002.
55. Guo, Y., “Application of the Ffowcs-Williams/Hawkings Equation to Two-
Dimensional Problems,” Journal of Fluid Mechanics, Vol. 403, Jan. 2000,
pp. 201–221.
56. Lockard, D. “An Efficient, Two-Dimensional Implementation of the Two-
Dimensional Ffowcs Williams Hawkings Equation,” Journal of Sound and
Vibration, Vol. 229, No. 4, 2000, pp. 897–911.
57. Amiet, R. K., “Refraction of Sound by a Shear Layer,” Journal of Sound and
Vibration, Vol. 58, No. 4, pp. 467–482.
58. Pilon, A. R., and Lyrintzis, A. S., “Refraction Corrections for the Modified
Kirchhoff Method,” AIAA paper No. 97-1654 presented at the 3rd AIAA/CEAS
Aeroacoustics Meeting, Atlanta, GA, May 1997; also AIAA Journal of Aircraft,
Vol. 35, No. 4, Jul.-Aug. 1998, pp. 661–664.
59. Freund, J. B., Lele, S. K., and Moin, P., “Calculation of the Radiated Sound Field
Using an Open Kirchhoff Surface,” AIAA Journal, Vol. 34, No. 5, May 1996,
pp. 909–916.
60. Brentner, K. S., “Numerical Algorithms for Acoustic Integrals with Examples for
Rotor Noise Prediction,” AIAA Journal, Vol. 35, No. 4, April 1997, pp. 625–630.
61. Singer, B., Lockard, D., and Brentner, K. S., “Computational Aeroacoustics
Analysis of Slat Trailing Edge Flow,” AIIA paper 99-1802 5th AIAA/CEAS
Aeroacoustics Conference, Bellevue, WA, May, 1999.
122 Surface integral methods in computational aeroacoustics

62. Meadows, K. R., and Atkins, H. R., “Towards a Highly Accurate Implementation
of the Kirchhoff Approach for Computational Aeroacoustics,” IMACS Journal of
Computational Acoustics, Vol. 4, No. 2, (1996), pp. 225–241.
63. Strawn, R. C., Biswas, R., and Lyrintzis, A. S., “Helicopter Noise Predictions
Using Kirchhoff Methods”, Proceedings of the 51st AHS Annual Forum, Fort
Worth TX, May 1995, Vol. I, pp. 495–508; also IMACS Journal of Computational
Acoustics, Vol. 4, No. 3, Sept. 1996, pp. 321–339.
64. Lyrintzis, A. S., Koutsavdis, E. K., Berezin, C., Visintainer, J., and Pollack, M.,
“An Evaluation of a Rotational Kirchhoff Methodology,” Journal of the American
Helicopter Society, Vol. 43, No. 1, Jan. 1998, pp. 48–57.
65. Polacsek, C., Prieur, J., “High-Speed Impulsive Noise Calculations in Hover
and Forward Flight Using a Kirchhoff Formulation,” CEAS/AIAA paper 95-138
Proceedings of the 1st Joint CEAS/AIAA Aeroacoustics Conference (16th AIAA
Aeroacoustics Conference), Munich Germany, June 1995, Vol. II, pp. 973–978.
66. Wissink, A. M., Lyrintzis, A. S., Strawn, R. C., Oliker, L., and Biswas, R.,
“Efficient Helicopter Aerodynamic and Aeroacoustic Predictions on Parallel
Computers,” AIAA paper No. 96-0153, presented at the AIAA 34th Aerospace
Science Meeting, Reno, NV, Jan. 1996.
67. Strawn, R. C., Oliker, L., and Biswas, R., “New Computational Methods for the
Prediction and Analysis of Helicopter Noise,” AIAA Journal of Aircraft, Vol. 34,
No. 5, pp. 665–672.
68. Long, L. N., and Brentner, K. S., “Self-Scheduling Parallel Methods for Multiple
Serial Codes with Applications to Wopwop, AIAA paper 2000-0346, presented at
the 38th AIAA Aerospace Sciences Meeting, Reno, NV, Jan. 1996.
69. Glegg, S. A. L., “The De-Dopplerization of Acoustic Signals Using Digital
Filters,” Journal of Sound and Vibration, Vol. 116, No. 2, 1987, pp. 384–387.
70. Ozyoruk, Y., and Long, L. N., “Computation of Sound Radiating from Engine
Inlets,” CEAS/AIAA paper 95-063 Proceedings of the 1st Joint CEAS/AIAA
Aeroacoustics Conference (16th AIAA Aeroacoustics Conference), Munich
Germany, June 1995; also AIAA Journal Vol. 34, No. 5, May 1996, pp. 894–901.
71. Ozyoruk, Y., and Long, L. N., “Multigrid Acceleration of a High Resolution
Computational Aeroacoustics Scheme,” AIAA Journal, Vol. 35, No. 3, March
1997, 428–433.
72. Ozyoruk, Y., and Long, L. N., “A New Efficient Algorithm for Computational
Aeroacoustics on Parallel Computers,” Journal of Computational Physics, Vol.
125, 1996, pp. 135–149.
73. Lyrintzis, A. S., and Xue, Y., “Towards a Versatile Kirchhoff’s Method Code,”
AIAA Journal, vol. 35, No. 1, Jan. 1997, pp. 198–200.
74. Rahier, G., and Prier, J., “An Efficient Kirchhoff Integration Method for Rotor
Noise Prediction Starting Indifferently from Subsonically or Supersonically
Rotating Meshes,” Proceedings of the 53rd AHS Annual Forum, Vol. 1, Virginia
Beach, VA, Apr. 1997, pp. 697–707.
aeroacoustics volume 2 á number 2 á 2003 123

75. Algermissen, G., and Wagner, S., “Computation of Helicopter High-Speed


Impulsive Noise by an Enhanced Kirchhoff Method,” Proceedings of the AHS
Technical Specialists’ Meeting for Rotorcraft Acoustics and Aerodynamics,
Williamsburg, VA, Oct. 1997.
76. Delrieux, Y., Prieur, J., Rahier, G., and Drousie, G., “A New Implementation of
Aeroacoustic Integral Method for Supersonic Deformable Control Surfaces,”
AIAA-2003-3201 presented at the 9th AIAA/CEAS Aeroacoustics Conference,
Hilton Head, SC, May 2003.
77. Kim, S., Dai, Y., Koutsavdis, E., Sovani, S., Kadam, N., and Ravuri, K., “A
Versatile Implementation of Acoustic Analogy Based Noise Prediction Method
in a General-Purpose CFD Code,” AIAA-2003-3202 presented at the 9th
AIAA/CEAS Aeroacoustics Conference, Hilton Head, SC, May 2003.
78. Brentner, K. S. “A New Algorithm for Computing Acoustic Integrals,” 14th
IMACS World Congress, Atlanta, GA, July 1994.
79. Forsyth, D. W., and Korkan, K. D., “Computational Aeroacoustics of Propeller
Noise in the Near- and the Far-field,” AIAA paper 87-0254, AIAA 25th Aerospace
Science Meeting, Reno, NV Jan. 1987.
80. Strawn, R. C., and Biswas, R., “Computation of Helicopter Rotor Noise in
Forward Flight,” Journal of the American Helicopter Society, Vol. 40, No. 3, July
1995, pp. 66–72.
81. Baeder, J. D., Gallman, J. M., and Yung, Y., “A Computational study of the
Aeroacoustics of Rotor in Hover,” Journal of the American Helicopter Society,
Vol. 42, No. 1, Jan. 1997, pp. 39–53.
82. Xue, Y., and Lyrintzis, A. S., “Rotating Kirchhoff Method for Three-Dimensional
Transonic Blade-Vortex Interaction Hover Noise,” AIAA Journal, Vol. 32, No. 7,
Jul. 1994, pp. 1350–1359.
83. Morgans, A., and Dowling, A.,“The Aeroacoustics of Transonic Helicopter
Blades,” AIAA paper 2002-2545 presented at the 8th AIAA/CEAS Aeroacoustics
Conference, Breckenridge, CO, June 2002.
84. Srinivasan, G. R., Baeder, J. D., Obayashi, S., and McCroskey, W. J., “Flowfield
of a Lifting Rotor in Hover - A Navier-Stokes Simulation,” AIAA Journal, Vol. 30,
No. 10, Oct. 1992, pp. 2371–2378.
85. Srinivasan, G. R., and Baeder, J. D., “TURNS: A Free-Wake Euler/Navier-Stokes
Numerical Method for Helicopter Rotors,” AIAA Journal, Vol. 31, No. 5, May
1993, pp. 959–962.
86. Buning, P. G., Jesperson, D. C., Pulliam, T. H., Chan, W. M., Stotnick, J. P., Krist,
S. E., and Renze, K. J., “OVERFLOW User’s manual Version 1.8g,” NASA
Langley Research Center, March 1999.
87. Strawn, R. C., and Caradonna, F. X., “Conservative Full Potential Model for
Unsteady Transonic Rotor Flows,” AIAA Journal, Vol. 25, No. 2, Feb. 1987,
pp. 193–198.
124 Surface integral methods in computational aeroacoustics

88. Caradonna, F. X., Strawn, R. C., and Bridgeman, J. O., “An Experimental and
Computational Study of Blade-Vortex Interactions” Vertica, Vol. 12, No. 4, 1988,
pp. 315–327.
89. Berezin, C., Pollack, M., Visintainer, J., Lyrintzis, A., and Koutsavdis, E.,
“Development and Practical Application of the Rotating Kirchhoff Method for the
Prediction of HSI and BVI Noise,” Proceedings of the AHS Technical Specialists’
Meeting for Rotorcraft Acoustics and Aerodynamics, Williamsburg, VA, Oct. 1997.
90. Hu, Z. W., Morfey, C. L., and Sandham, N. D., “Sound Radiation from a Subsonic
Turbulent Plane Jet,” AIAA paper No. 2002-2421 presented at the 8th
AIAA/CEAS Aeroacoustics Conference, Breckenridge, CO, June 2002.
91. Hardin, J. C., Ristorcelli, J. R. and Tam, C. K. W. editors. “ICASE/LARC
Workshop on Benchmark Problems in Computational Aeroacoustics,” NASA
Conference Publication 3300, NASA Langley, May 1995.
92. Singer, B., Lockard, D., Brentner, K. S., and Lilley, G. M., “Simulation of
Acoustic Scattering from a Trailing Edge,” Journal of Sound and Vibration, Vol.
230, No. 3, 2000, pp. 541–560.
93. Pierce, A., Acoustics – An Introduction to Its Physical Principles and
Applications, Acoustical Society of America, 1989.
94. Jaeger, S., and Korkan, K. D., “On the Prediction of Far-Field Computational
Aeroacoustics of Advanced Propellers,” AIAA paper 90-3996, AIAA 13th
Aeroacoustics Conference, Oct. 1990.
95. Lyrintzis, A. S., “The Use of Kirchhoff’s Method in Rotorcraft Aeroacoustics,” Paper
No. 34, presented at the 75th AGARD Fluid Dynamics Panel Meeting and
Symposium on Aerodynamics and Aeroacoustics of Rotorcraft, Berlin, Germany,
Oct. 1994; AGARD Conference Proceedings, No. 552, Aug. 1995, pp. 34-1–34-16.
96. George, A. R., and Lyrintzis, A. S., “Acoustics of Transonic Blade-Vortex
Interactions,” AIAA Journal, Vol. 26, No. 7, Jul. 1988, pp. 769–776.
97. Lyrintzis, A. S., and George, A. R., “Far-Field Noise of Transonic Blade-Vortex
Interactions,” American Helicopter Society Journal, Vol. 34, No. 3, July 1989,
pp. 30–39.
98. Lyrintzis, A. S., and Xue, Y., “A Study of the Noise Mechanisms of Transonic Blade-
Vortex Interactions,” AIAA Journal, Vol. 29, No. 10, Oct. 1991, pp. 1562–1572.
99. Xue, Y., and Lyrintzis, A. S., “Transonic Blade-Vortex Interactions: Noise Reduction
Techniques,” AIAA Journal of Aircraft, Vol. 30, No. 3, May-June, 1993, pp. 408–411.
100. Lyrintzis, A. S., Lee, J., and Xue, Y., “Mechanisms and Directivity of Unsteady
Transonic Flow Noise” presented at the International Symposium on Flow
Induced Vibration and Noise III, Vol. 3: Flow-Structure and Flow-Sound
Interactions, eds. Farabee, T. M., Paidoussis, M. P., pp. 85–113, ASME Winter
Annual Meeting, Anaheim, CA Nov. 1992; also ASME Journal of Fluids
Engineering Vol. 116, No. 3, Sept. 1994, pp. 649–652.
101. Lin, S-Y and Chin, Y-S, “Numerical Study on Reduction of Transonic Blade-
Vortex Interaction Noise,” CEAS/AIAA paper 95-049 Proceedings of the 1st Joint
aeroacoustics volume 2 á number 2 á 2003 125

CEAS/AIAA Aeroacoustics Conference (16th AIAA Aeroacoustics Conference),


Munich Germany, June 1995.
102. Lin, S-Y and Chen, Y-F, “Numerical Study of Head on Blade-Vortex Interaction
Noise,” AIAA paper No. 97-0287 presented at the 35th Aerospace Science
Meeting, Reno, NV Jan. 1997.
103. Strawn, R. C., Garceau, M., and Biswas, R., “Unstructured Adaptive Mesh
Computations of Rotorcraft High-Speed Impulsive Noise,” AIAA Journal of
Aircraft, Vol. 32, No. 4, July-Aug. 1995, pp. 754–760.
104. Ahmad, J. U., Duque, E. P. N., and Strawn, R. C., “Computations of Rotorcraft
Aeroacoustics with a Navier Stokes/Kirchhoff Method,” paper 51, 22nd European
Rotorcraft Forum, Brighton, UK, Sept. 1996.
105. Lyrintzis, A. S., Xue, Y., and Kilaras, M. S., “The Use of a Rotating Kirchhoff
Formulation for High-Speed Impulsive Noise,” AIAA Paper 94-0463, presented
at the 32nd AIAA Aerospace Sciences Conference, Reno, NV, Jan. 1994.
106. Lyrintzis, A. S., Kilaras, M. S., and Xue, Y., “Transonic 3-D BVI Noise Using a
Rotating Kirchhoff Formulation for Advancing Rotors,” Proceedings of the 50th
AHS Annual Forum, Washington, DC, May 1994, Vol. I, pp. 115–127.
107. Lyrintzis, A. S., and Koutsavdis, E. K., “Rotorcraft Impulsive Noise Prediction
Using a Rotating Kirchhoff Formulation,” AIAA Journal of Aircraft, Vol. 33,
No. 6, Nov.-Dec. 1996, pp. 1054–1061.
108. Schmitz, F. H., Boxwell, D. A., Splettstoesser, W. R., and Schultz, K. J., “Model-
Rotor High-Speed Impulsive Noise: Full-Scale Comparisons and Parametric
Variations,” Vertica, Vol. 8, No. 4, 1984, pp. 395–422.
109. Kuntz, M., Lohmann, D., Pahlke, K., “Comparisons of Rotor Noise Predictions at
DLR Obtained by a Lifting Surface Method and Euler Solutions Using Kirchhoff
Equation,” CEAS/AIAA paper 95-136 Proceedings of the 1st Joint CEAS/AIAA
Aeroacoustics Conference (16th AIAA Aeroacoustics Conference), Munich
Germany, June 1995, Vol. II, pp. 949–962.
110. Kuntz, M., “Rotor Noise Predictions in Hover and Forward Flight Using Different
Aeroacoustic Methods,” AIAA paper No. 96-1695, presented at the 2nd
AIAA/CEAS Aeroacoustics Meeting, (17th AIAA Aeroacoustics Meeting) State
College, PA, May 1996.
111. Lyrintzis, A. S., Koutsavdis, E. K., and Strawn, R. C., “A Comparison of
Computational Aeroacoustic Prediction Methods,” American Helicopter Society
Journal, Vol. 42, No. 1, Jan. 1997, pp. 54–57.
112. Brentner, K. S., Lyrintzis, A. S., and Koutsavdis, E. K., “A Comparison of
Computational Aeroacoustic Prediction Methods for Transonic Rotor Noise,”
AIAA Journal of Aircraft, Vol. 34, No. 4, Jul.-Aug. 1997, pp. 531–538.
113. Purcell, T. W., “CFD and Transonic Helicopter Sound,” Paper No. 2, 14th
European Rotorcraft Forum, Sept. 1988.
114. Purcell, T. W., “A Prediction of High-Speed Rotor Noise,” AIAA 89-1130, AIAA
12th Aeroacoustics Conference, San Antonio, TX, 1989.
126 Surface integral methods in computational aeroacoustics

115. Lyrintzis, A. S., Koutsavdis, E. K., and Pilon, A. R., “An Extended Kirchhoff Method
for Rotorcraft Impulsive Noise,” Proceedings of the AHS Technical Specialists’
Meeting for Rotorcraft Acoustics and Aerodynamics, Williamsburg, VA, Oct. 1997.
116. Davis, C. M., and Atassi, H. M., “The Far-Field Acoustic Pressure of an Airfoil in
Nonuniform Subsonic Flows,” presented at the Symposium of Flow Noise
Modeling, Measurement and Control, NCA-Vol. 11/ FED-Vol. 130, pp. 107–117,
ASME Winter Annual Meeting, Atlanta, GA, Dec. 1991.
117. Atassi, H. M., Dusey, M., and Davis, C. M., “Acoustic Radiation from a Thin Airfoil
in Nonuniform Subsonic Flows,” AIAA Journal, Vol. 31, No. 1, Jan. 1993, pp. 12–19.
118. Patrick, S. M., Davis, C. M., and Atassi, H., “Acoustic Radiation from a Lifting
Airfoil in Nonuniform Subsonic Flows” in Computational Aero- and Hydro-
Acoustics, FED Vol. 147, eds: Mankbadi, R. R., Lyrintzis, A. S., Baysal,
O., Povinelli, L. A., and Hussaini, M. Y., pp. 41–46, ASME Fluids Engineering
Conference, Washington, DC, June 1993.
119. Scott, J. R., and Atassi, H. M., “Numerical Solution of the Linearized Euler
Equations for Unsteady Vortical Flows Around Lifting Airfoils,” AIAA paper No
90-0064, AIAA 28th Aerospace Science Meeting, Reno, NV.
120. Scott, J. R., and Atassi, H. M., “A Finite-Difference Frequency Domain,
Numerical Scheme for the Solution of the Gust Response Problem,” Journal of
Computational Physics,Vol. 119, pp. 75–93.
121. Fang J., and Atassi, H. M., “Direct Calculation of Sound Radiated from a Loaded
Cascade in a Gust,” in Computational Aero- and Hydro-Acoustics, FED Vol. 147,
eds: Mankbadi, R. R., Lyrintzis, A. S., Baysal, O., Povinelli, L. A., and Hussaini,
M. Y., pp. 111–116, ASME Fluids Engineering Conference, Washington, DC,
June 1993.
122. Zhang X., Chen, X. X., Morfey, C. L., Nelson, P. A., “Computation of Sound
Radiation of an Exhaust Duct,” presented at the CEAS Workshop “From CFD to
CAA,” Nov. 2002, Athens Greece.
123. Soh, W. Y., 1994, “Unsteady Jet Flow Computation-Towards Noise Prediction,”
AIAA paper 94-0138, AIAA 32nd Aerospace Science Meeting, Reno, NV Jan. 1994.
124. Mitchell, B. E., Lele, S. K., and Moin, P., “Direct Computation of the Sound
Generated by Vortex Pairing in an Axisymmetric Jet,” Journal of Fluid
Mechanics, Vol. 383, 1999, pp. 113–142.
125. Zhao, W., Frankel, S. H., and Mongeau, L., “Large eddy simulations of sound
radiation from subsonic turbulent jets,” AIAA Journal, Vol. 39, No. 8, Aug. 2001,
pp. 1469–1477.
126. Billson, M., Eriksson, L., and Davidson, L., “Jet Noise Prediction Using
Stochastic Turbulence Modeling,” AIAA-2003-3282 presented at the 9th
AIAA/CEAS Aeroacoustics Conference, Hilton Head, SC, May 2003.
127. Chyczewski, T. S., and Long, L. N., “Numerical Prediction of the Noise Produced
by a Perfectly Expanded Rectangular Jet,” AIAA Paper 96-1730, 2nd
AIAA/CEAS Aeroacoustics Conference, State College, PA, May 1996.
aeroacoustics volume 2 á number 2 á 2003 127

128. Morris, P. J., Long, L. N., Scheidegger, T., Wang, Q., and Pilon, A. R., “High
Speed Jet Noise Simulations,” AIAA Paper No. 98-2290 presented at the 4th
AIAA/CEAS Aeroacoustics Conference, Toulouse, France, June, 1998.
129. Gamet l., and Estivalezes, J. L., “Application of Large-Eddy Simulations and
Kirchhoff Method to Jet Noise Prediction,” AIAA Journal, Vol. 36, No. 12, Dec.
1998, pp. 2170–2178.
130. Choi, D., Barber, T. J., Chiappetta, L. M., and Nishimura, M., “Large Eddy
Simulations of high-Reynolds number jet flows,” AIAA Paper No. 99-0230
presented at the AIAA, 37th Aerospace Sciences Meeting and Exhibit, Reno, NV,
Jan. 1999.
131. Kandula, M., and Caimi, R., “Simulation of Jet Noise with Overflow CFD
Code and Kirchhoff Surface Integral,” AIAA paper No. 2002-2602, presented
at the 8th AIAA/CEAS Aeroacoustics Conference, Breckenridge, CO, June
2002.
132. Balakumar, P., “Prediction of Supersonic Jet Noise,” AIAA paper No. 98-1057,
presented at the 35th Aerospace Science Meeting, Reno, NV, Jan. 1998.
133. Yen, C-C., and Messersmith, N. L., “The Use of Compressible Parabolized
Stability Equations for Prediction of Jet Instabilities and Noise,” AIAA
Paper 99-1859 5th AIAA/CEAS Aeroacoustics Conference, Bellevue, WA, May
1999.
134. Shih, S. H., Hixon, D. R., Mankbadi, R. R., Pilon, A. R., and Lyrintzis, A. S.,
“Evaluation of Far-Field Jet Noise Prediction Methods,” AIAA paper No. 97-
0282, presented at the 35th Aerospace Science Meeting, Reno, NV, Jan. 1997.
135. Hu, Z. W., Morfey, C. L., and Sandham, N. D., “Large Eddy Simulation of Plane
Jet Sound Radiation,” AIAA paper No. 2003-3166 presented at the 9th
AIAA/CEAS Aeroacoustics Conference, Hilton Head, SC, May 2003.
136. Morris, P. J., Scheidegger, T., and Long L. N., “Jet Noise Simulations for Circular
Nozzles,” AIAA paper No. 2000-2080, presented at the 6th AIAA/CEAS
Aeroacoustics Conference, Lahaina, HA, June 2000.
137. Boluriaan, S., Morris, P. J., Long L. N., and Scheidegger, T., “High Speed
Jet Noise Simulations for Noncircular Jets,” AIAA paper 2001-2147 presented
at the 7th AIAA/CEAS Aeroacoustics Conference, Maastricht, Netherlands,
May 2001.
138. Uzun, A., Blaisdell, G. A., and Lyrintzis, A. S., “3-D Large Eddy Simulation for
Jet Aeroacoustics,” AIAA paper 2003-3322 presented at the 9th AIAA/CEAS
Aeroacoustics Conference, Hilton Head, SC, May 2003.
139. Rahier G., Prieur J., Vuillot, F., Lupoglazoff, N., Biancherin A., “Investigation of
Integral Surface Formulations for Acoustic Predictions of Hot Jets Starting From
Unsteady Aerodynamic Simulations,” AIAA paper 2003-3164 presented at the 9th
AIAA/CEAS Aeroacoustics Conference, Hilton Head, SC, May 2003.
140. Gliebe, P. R., “High Velocity Jet Noise Source Location and Reduction,” Task 2,
FAA-RD-76-79-II, May 1978.
128 Surface integral methods in computational aeroacoustics

141. Wundrow, D. W., Khavaran, A., “On the Applicability of High-Frequency


Approximations to Lilley’s Equation,” NASA/CR-2003-212089, Jan. 2003.
142. Sarigul-Klijn, N., Dietz, D., Karnopp, D., and Dummer, J., “A Computational
Aeroacoustic Method for Near and Far Field Vehicle Noise Predictions,” AIAA paper
2001-0513, presented at the 39th Aerospace Science Meeting, Reno, NV, Jan. 2001.
143. Morris, P. J., Private Communication, July 2003.
144. Curle, N., “The Influence of Solid Boundaries Upon Aerodynamical Sound,”
Proc. Roy. Soc. London, A231, 1955, pp. 505–514.
145. Wang, M., Lele, S. K., and Moin P., “Computation of Quadrupole Noise Using
Acoustic Analogy,” AIAA Journal, Vol. 34, No. 11, Nov. 1996.

You might also like