You are on page 1of 26

Fluid Dynamics Research

You may also like


- A Reliable and Sensitive Voltammetric
Recent advances in computational aeroacoustics Determination of Mo(VI) at the In Situ
Renovated Bismuth Bulk Annular Band
Electrode
To cite this article: Christopher K W Tam 2006 Fluid Dyn. Res. 38 591 Krystian Wgiel, Magorzata Grabarczyk,
Wadysaw W. Kubiak et al.

- Oxidative Polymerization of Polyaniline


(PANI) Colloids with Different Oxidizing
Agents
View the article online for updates and enhancements. Ashwini B. Rohom, Priyanka U. Londhe
and Nandu B. Chaure

- Prebiotic Cytosine Synthesis from Urea in


Interstellar Space: A Computational
Mechanistic Study
Joong Chul Choe

This content was downloaded from IP address 1.47.70.223 on 29/08/2022 at 07:07


FLUID DYNAMICS RESEARCH is now available from the IOP Publishing website http://www.iop.org/journals/FDR

Fluid Dynamics Research 38 (2006) 591 – 615

Recent advances in computational aeroacoustics


Christopher K.W. Tam∗
Department of Mathematics, Florida State University, Tallahassee, FL 32306-4510, USA
Received 11 September 2003; received in revised form 28 December 2005; accepted 8 March 2006

Communicated by K. Ishii

Abstract
There are intrinsic differences between the characteristics and the objectives of aeroacoustics problems and typ-
ical fluid dynamics problems. These differences are not addressed by computational fluid dynamics (CFD). They
form some of the major challenges to computational aeroacoustics (CAA). Recent advances in CAA methodologies
to meet those challenges are reviewed. They include the development of high resolution CAA methods, artificial
selective damping and high quality numerical boundary conditions. One unique development in CAA that has no
counterpart in CFD is the wave number analysis approach. Wave number analysis not only can yield an absolute
error incurred in the use of a discretized computational method but also offers a way to develop optimized compu-
tation schemes. It is indispensable in analyzing dispersion and dissipation errors associated with wave propagation
computations. To illustrate the impact of CAA on aeroacoustics, an example of CAA application will be briefly
discussed. As CAA application expands in scope, new challenges to CAA emerge. Some of the more pressing
challenges that may well chart the future development of CAA methodology are elaborated.
© 2006 The Japan Society of Fluid Mechanics and Elsevier B.V. All rights reserved.
PACS: 43.20; 43.25; 43.28; 2.60; 2.70

Keywords: Computational aeroacoustics; High resolution scheme; Numerical dispersion and dissipation; Wave number
analysis; Artificial selective damping; Radiation and outflow boundary conditions

1. Introduction

Aeroacoustic problems are by nature very different from standard aerodynamics and fluid mechanics
problems. Before discussing how to solve aeroacoustics problems numerically or simulate them com-
∗ Tel.: +1 850 644 2455; fax: +1 850 644 4053.
E-mail address: tam@math.fsu.edu.

0169-5983/$30.00 © 2006 The Japan Society of Fluid Mechanics and Elsevier B.V. All rights reserved.
doi:10.1016/j.fluiddyn.2006.03.006
592 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

putationally, an approach generally referred to as computational aeroacoustics (CAA), it is important to


recognize and to have a good understanding of these differences. These differences pose a number of
major challenges to CAA. A few of the important computational challenges are listed below.

(a) Aeroacoustics problems, by definition, are time dependent, whereas aerodynamics and fluid mechanics
problems are generally time independent or involve only low frequency unsteadiness.
(b) Aeroacoustics problems typically involve frequency range that spreads over a wide bandwidth. Numer-
ical resolution of the high frequency waves with extremely short wavelengths becomes a formidable
obstacle to accurate numerical simulation.
(c) Acoustic waves usually have small amplitudes. They are very small compared to the mean flow.
Oftentimes, the sound intensity is five to six orders smaller. To compute sound waves accurately, a
numerical scheme must have extremely low numerical noise.
(d) In most aeroacoustics problems, interest is in the sound waves radiating to the far field. This requires
a solution that is uniformly valid from the source region all the way to the measurement point at
many acoustic wavelengths away. Because of the long propagation distance, CAA schemes must
have minimal numerical dispersion and dissipation. Also, it should propagate the waves at the correct
wave speeds and is isotropic irrespective of the orientation of the computation mesh.
(e) In general, flow disturbances in aerodynamics or fluid mechanics problems tend to decay very fast
away from a body or their source of generation. Acoustic waves, on the other hand, decay very slowly
and actually reach the boundaries of a finite computation domain. To avoid the reflection of outgoing
sound waves back into the computation domain, and thus contaminates the numerical solution, radi-
ation boundary conditions must be imposed at the artificial exterior boundaries to assist the waves to
exit smoothly. For standard computational fluid dynamics (CFD) problems, such boundary conditions
are usually not required.
(f) Aeroacoustics problems are archetypical examples of multiple-scales problems. The length scale of
the acoustic source is usually very different from the acoustic wavelength. That is, the length scale
of the source region and that of the acoustic far field region can be vastly different. CAA methods
must be designed to deal with problems with greatly different length scales in different parts of the
computational domain.

It must be acknowledged that CFD has been very successful in solving fluid and aerodynamics prob-
lems. CFD methods are generally designed for computing time independent solutions. Because of the
tremendous success of CFD, it is tempting to use these methods to solve aeroacoustics problems as well.
In the past, there have been a number of attempts to do just that. However, the results have proven to be
quite discouraging. For example, Hsi and Perie (1977) tried to use a commercial CFD code RADIOSS to
solve the sound scattering problems of the Second CAA Workshop on Benchmark Problems. The results
were disastrous. The computed results were highly dispersive and differed significantly from the exact
solutions.
As elaborated above, it should be clear that the nature of aeroacoustics problems are substantially
different from those of traditional fluid dynamics and aerodynamics problems. To be able to compute or
simulate aeroacoustics problems accurately and efficiently, standard CFD schemes, designed for appli-
cations to fluid problems, are generally not adequate. For this reason, there is a need for an independent
development of CAA. This was what happened in the last 10 years. During this period, great advances have
been made both in the development of CAA methodology and in their applications to real world problems.
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 593

There have been a number of review articles written on CAA during recent years by Tam (1995), Lele
(1997), Wells and Renaut (1997) and others. The objective of this paper is to review some of the more
recent advances in CAA. Specifically, the construction of optimized high resolution CAA schemes will
be discussed. The origin of numerical dispersion and dissipation, two principal wave propagation com-
putation errors, will be examined and analyzed through wave number analysis based on Fourier-Laplace
transforms. Any CAA algorithm must include some form of artificial selective damping or filtering. They
are to eliminate spurious short waves and to promote numerical stability. The construction and analysis
of high quality artificial numerical damping stencils will be reviewed. A necessary component of a CAA
algorithm is numerical boundary conditions. In general, there are two types of numerical boundary con-
ditions in CAA. They are asymptotic boundary conditions and absorbing boundary conditions. In this
paper, a brief discussion of asymptotic radiation and outflow boundary conditions is presented. CAA
is not methodology alone. A very important part of CAA is application. Here, as an example of CAA
application, direct numerical simulation (DNS) of acoustic liners is discuss. For this problem, emphasis
is on the use of CAA to investigate the mechanisms by which acoustic energy is dissipated by a resonant
acoustic liner. To conclude this work, some of the pressing challenges confronting CAA at the present
time are examined.

2. Optimized high resolution schemes

One of the basic needs of CAA is schemes that can resolve high frequency short waves using only a
few mesh points per wave length. Standard second-order CFD schemes often require 18–25 mesh points
per wavelength. This large number of mesh points makes the scheme impractical for use as a design
tool. The need to use minimum number of mesh points to resolve a wave led to the development of
high resolution large stencil CAA schemes. Currently the compact scheme reformulated by Lele (1992)
and the dispersion-relation-preserving (DRP) scheme of Tam and Webb (1993) are the most widely used
schemes in CAA. Here, we will discuss the DRP scheme as the optimization procedure of this scheme
has found application in many new schemes.
Consider a uniform mesh of spacing x as shown in Fig. 1. We will approximate the spatial derivative
at x = x by a finite difference quotient with N points to the left and M points to the right of , i.e.
 
1 
M
ju

aj u+j . (1)
jx  x
j =−N

Traditionally, the stencil coefficients aj are found by expanding the right side of (1) as Taylor series in x
and then equating coefficients of the same powers of x. This truncated Taylor series method, popular
as it is, does not provide a way to ascertain the error of approximation in quantitative terms. Nor does

Fig. 1. A uniform mesh of spacing x.


594 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

it offer any information on propagation errors when the finite difference approximation is used to solve
wave propagation problems.
Tam and Webb (1993) considered the finite difference approximation (1) from the wave number space
point of view. In their approach, x was taken as finite size, not necessarily small. They regarded the
coefficients aj to be free parameters that were to be determined so that (1) was an optimized approximation.
Following Tam and Webb, we may consider Eq. (1) to be a relationship between the values of a function
on a set of discrete mesh points on the x-axis (see Fig. 1). We may generalize this relationship and regard
it to hold true not just for the set of points labeled by the index  but to hold true for any set of points
spaced at x apart on the x-axis. This generalizes (1) to

1 
M
ju
(x)
aj u(x + j x). (2)
jx x
j =−N

In (2) x is an arbitrary point on the x-axis. By setting x = x in (2), Eq. (1) is recovered. Eq. (2) is a finite
difference relation of a continuous variable.
Let f˜() be the Fourier transform of f (x). The function f (x) and its transform f˜() are related by
 ∞  ∞
1
f˜() = f (x) e −i  x
dx, f (x) = f˜()eix d, (3)
2 −∞ −∞
where  is the wave number or the Fourier transform variable. By taking the Fourier transform of (2), it
is straightforward to find, with the help of the Shifting Theorem
iũ
i¯ũ, (4)
where
−i 
M
¯ = aj eij x . (5)
x
j =−N

In the case of a symmetric stencil, i.e., N = M and setting a−j = −aj , (5) becomes

2 
N
¯ = aj sin(x). (6)
x
j =1

The  on the left side of (4) is the wave number arising from the transform on the x derivative. The ¯ on
the right side may, therefore, be interpreted as the wave number of the finite difference scheme. It is to
be noted from (6) that for central difference approximation ¯ is real when  is real. On the other hand for
nonsymmetric difference approximation ¯ is complex even when  is real.
Once the aj ’s are known, the relationship between , the true wave number, and ¯ , the wave number of
the finite difference scheme, may be calculated by (5) or (6). High resolution schemes can be constructed
by choosing the stencil coefficients aj to minimize the square of the absolute value of the difference
between ¯ x and x over a large band of wave number. That is, aj ’s are determined to minimize the
integral
 
E= |¯x − x|2 d(x). (7)
−
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 595

Tam and Webb considered central difference stencils. They imposed the extra constraint that the Taylor
series expansion of (1) is satisfied to order 2(N −1). With this extra constraint, there is one free parameter,
say a1 . This parameter is chosen so that E is a minimum or jE/ja1 = 0. For N = 3 and N = 7, they
recommended the following stencil coefficients:

a−j = −aj , a0 = 0.

N = 3,  = 1.1

a1 = 0.770882380518225552,

a2 = −0.166705904414580469,

a3 = 0.0208431427703117643.

N = 7,  = 1.8

a1 = 0.91942501110343045059277722885,

a2 = −0.35582959926835268755667642401,

a3 = 0.15251501608406492469104928679,

a4 = −0.059463040829715772666828596899,

a5 = 0.019010752709508298659849167988,

a6 = −0.0043808649297336481851137000907,

a7 = 0.00053896121868623384659692955878.

Fig. 2 shows the dependence of ¯ x on x over the interval −x  for the standard sixth order
central difference scheme, the 7-point stencil and the 15-point stencil DRP scheme of Tam and Webb. For
low wave numbers, ¯ x is nearly equal to x (¯x is identically equal to x on the 45◦ line of Fig. 2).
For instance, for the 7-point stencil DRP scheme, ¯ x differs from x by less than 0.1% in the interval
0  |x|  1.05. That is, this scheme incurs an error less than 0.1% when 6 mesh points per wavelength
is used in a computation.
For unsymmetric stencils (N not equal to M),¯x is complex for real x. It will be shown later that
the imaginary part of ¯ x is related to the numerical damping or growth of a wave in a computation.
For a stable algorithm, if c (c may be positive or negative) is the wave speed then the numerical stability
condition for simple wave is

c Im(¯x) < 0, − x . (8)

To take condition (8) into account, the stencil coefficients may be chosen to minimize the integral
   
F= |Re(¯x) − x| d(x) + 
2
|Im(¯x) +  sgn(c)e−(ln 2)((x−)/) |2 d(x), (9)
0 0
596 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

Fig. 2. The ¯ x versus x curves. .............. Sixth order central difference scheme, ——— 7-point DRP scheme, – – – – –
15-point DRP scheme.

where , ,  and  are adjustable parameters. An example of an unsymmetric 7-point stencil (N = 4,


M = 2) for right propagating wave is

a−4 = 1.6111145483497D − 02,

a−3 = −0.12586455863612,

a−2 = 0.47098894426146,

a−1 = −1.2808664336879,

a0 = 0.53364403960380,

a1 = 0.42401136509001,

a2 = −3.8024502114702D − 02.

Note: For unsymmetric stencils ¯ x(−x) = −¯∗ x(x) where ∗ denotes the complex conjugate.
Fig. 3 shows the dependence of the real part of ¯ x on x for the above (4,2) stencil. Fig. 4 shows
the imaginary part of ¯ x as a function of x. Condition (8) is satisfied for right propagating wave.
Recently, a number of optimized unsymmetric (upwind) schemes were proposed by Lockard et al. (1995),
Li (1997), Zhuang and Chen (1998, 2002), Zingg (2000), and others for use in CAA. Interested readers
should consult these references for details of their optimization criteria.
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 597

Fig. 3. The dependence of Re(¯x) on x for the (4,2) stencil.

Fig. 4. The dependence of Im(¯x) on x for the (4,2) stencil.

3. Numerical dispersion

Acoustics are governed by the compressible Navier–Stokes equations. Molecular viscosity, in most
case, is unimportant so that it is sufficient to use the Euler equations. To solve the Navier–Stokes or
Euler equations computationally, it is necessary to perform discretized approximation to the spatial and
598 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

temporal derivatives. Once discretized, the equations are no longer the same as those of the original
partial differential equations. The discretized equations have their own exact solutions. A central effort
of CAA is to understand the mathematical behavior of the exact solution of the discretized equations and
to quantify and minimize the error. Here, error is referred to as the difference between the solutions of
the original partial differential equations and the discretized system.
Invariably, the discretized equations behave mathematically like a dispersive wave system (Vichnevet-
sky and Bowles, 1982; Trefethen, 1982; Tam and Webb, 1993), even though the waves supported by
the original partial differential equations are nondispersive. This is a very important point and should be
clearly understood by all CAA investigators and users. As an illustration, let us consider the solution of
the simple convective equation
ju ju
+c =0 (10)
jt jx
and initial condition
t =0 u(x, 0) = (x).
To find the nature of the waves supported by this equation, we may perform a Fourier-Laplace transform
to this equation. The Laplace transform of a function g(t), denoted by g̃(), and its inverse are related by
 ∞ 
1
g̃() = g(t)eit dt, g(t) = g̃() e−it d,
2 −∞

where  is the angular frequency (Laplace transform variable). is a contour in the upper half -plane
parallel to the real axis, above all poles and singularities of g̃().
The Fourier-Laplace transform of (10) leads to,
iˆ ()
( − c)ũ = , (11)
2
where ũ is the Fourier-Laplace transform of u.ˆ () is the Fourier transform of initial condition u(x, 0).
Thus ũ is given by,
iˆ ()
ũ = . (12)
2  ( − c  )
The poles of (12) are found by setting the denominator to zero. This yields,
 = c. (13)
Eq. (13) gives the relationship between wave number and angular frequency. It is called the dispersion
relation. According to dispersive wave theory (Whitman, 1974), all wave propagation characteristics of
a partial differential equation are encoded in the dispersion relation. The group velocity or wave speed is
given by d/d. In the case of (10), by using dispersion relation (13), the group velocity is,
d
=c (14)
d
so that all wave components propagate with the same speed. That is to say, the waves supported by the
convective wave equation are nondispersive.
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 599

Now we may convert (10) into a discretized system on a uniform mesh as shown in Fig. 1 by first
approximating the x-derivative at the th mesh point by a standard central difference quotient or by the
DRP scheme (Tam and Webb, 1993) with a stencil of (2N + 1) points, i.e.,
 
1 
N
ju

aj u+j (aj = −a−j ). (15)
j x  x
j =−N

The semi-discretized form of (10) may be rewritten as a system of equations,


du
= K , (16a)
dt

−c 
N
K = aj u+j . (16b)
x
j =−N

Eq. (16a) is a finite difference differential equation involving integer variable . The equation can be
generalized to one with continuous variables (x, t) by assuming that the relations hold true for any set of
points at spacing x on the x-axis. The generalized equations are
ju
= K,
jt

c 
N
K =− aj u(x + j x, t). (17)
x
j =−N

Eq. (16) is a special case of (17). Now by taking the Fourier-Laplace transforms of (17), it is easy to
find
iˆ ()
ũ = . (18)
2( − c¯)
Eq. (18) differs from (12) only in that  is replaced by ¯ in the denominator. By inverting the Fourier
and Laplace transforms, the exact solution of the discretized (16) is found to be,
 ∞
i ˆ ()

u(x, t) = ei(x−t) d d. (19)
2 −∞  − c¯()
Dispersion relation (13) is now replaced by
 = c¯(). (20)
The integrand has a pole at  = c¯(). On deforming the inverse contour to pick up this pole in the
-plane, it is easy to find through the use of the Residue Theorem
 
∞ 
i((x/t)−)t 
u(x, t) = ˆ ()e
 d . (21)

−∞ =c¯()

Now for large t (with x/t fixed), the -integral of (21) may be evaluated by the method of stationary
phase (see e.g., Ablowitz and Fokas, 1997, Chapter 6). The stationary phase point  = s is given by the
600 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

Fig. 5. The d¯/d versus x curves. ........... Sixth order central difference scheme, ———– 7-point DRP scheme, – – – –
15-point DRP scheme.

zero of the derivative of the phase function


= ((x/t) − ()) with respect to . This leads to

d d¯()  x
=c  = , (22)
d d =s t
which yields s as a function of x/t. The asymptotic solution is
 1/2
2 )
u(x, t) ∼ 
ˆ (s )ei[s (x/t)−c¯(s )]t+i(/4)sgn(−c¯ ,
 (23)
t→∞ |c¯ (s )|t
where sgn(z) is the sign of the real part of z.
In wave propagation theory (Whitman, 1974) d/d in (22) is called the group velocity. Since x =
(d/d)t, it is the propagation speed of the component of the solution with wave number  in the x .t
diagram. Now  and  are related by (20) so that d/d = c(d¯/d). In general d¯/d is not a constant
equal to 1. Different wave number of the initial disturbances will propagate with a slightly different
speed. Because of the different propagation speed a wave packet will spread out or disperse in space as
it propagates. Even a small difference in group velocity can manifest into serious dispersion over a long
time or a long distance of propagation. This is the origin of numerical dispersion.
For the convective wave (10), (22) reveals that the variation in group velocity is caused by the variation
of the slope of the ¯ x versus x curve. Fig. 5 shows the d¯/d versus x curves for a number of
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 601

Fig. 6. The comparison between computed solutions and exact solution for the one-dimensional convective wave equation: (a)
sixth order central difference scheme, (b) 7-point DRP scheme, (c) 15-point DRP scheme.

schemes. Notice that the d¯/d curve for the 7-point stencil DRP scheme has a peak at x = 0.67 at
which d¯/d = 1.00276. That is, the wave component with x = 0.67 (wave with 9.38 mesh points
per wavelength) will propagate faster than the exact wave speed d¯/d = 1 by about 0.3%. Suppose in a
computation, the waves propagate over a distance of 400 mesh points. In this case, when the main part of
the wave packet reaches x = 400, the wave component with x = 0.67 will reach x = 401.1. Under these
circumstances, numerical dispersion would just be noticeable as wave dispersion exceeds 1 mesh point.
On the other hand, if the standard sixth order scheme is used instead, severe numerical dispersion will
result when the main wave reaches x = 400. Waves in the wave number range of x > 0.6, having group
velocity less than 1.0 would form trailing waves. However, the d¯/d curve for the 15-point stencil DRP
scheme is practically equal to1 for an extended range of x. For x < 1.73, the group velocity differs
from 1 by no more than 0.1%. Thus, if the wave packet of the solution contains only waves with wave
number x < 1.73, then there will be no observable numerical dispersion even after the wave packet
propagates over a distance of up to 1000 mesh points.
As an example, Fig. 6 shows the computed solution of the convective wave equation (10) with initial
condition in the form of a Gaussian with a half-width of 3 mesh spacings,

u(x, 0) = e−(ln 2)(x/3x)


2
(24)
602 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

using the standard sixth order scheme, the 7-point stencil and the 15-point stencil DRP scheme. It is easy
to see that there is significant numerical dispersion (with extensive trailing waves) when computed by the
standard sixth order scheme. The 7-point stencil DRP scheme is better. Both schemes, having the same
stencil size, require the same amount of computation. The solution by the 15-point stencil is good. For
all intents and purposes, the solution is identical numerically to the exact solution.
It is worthwhile to point out that numerical dispersion is the result of the variation of the group velocity
of a numerical scheme. Unfortunately, in many textbooks it is linked erroneously to the phase velocity
of the computation scheme. Phase velocity and group velocity are not the same. In fact, they can have
opposite signs so that phase velocity and group velocity can propagate in opposite directions. From
Fig. 5, it is easy to see that d¯/d for the standard sixth order central difference scheme as well as the
7-point stencil DRP scheme is negative for x > 2.0. In other words, for short waves with x > 2.0,
the phase velocity is positive but the waves propagate backward because the group velocity is negative.
We would like to emphasize again that for spatial resolution, it is important to require ¯ x of a scheme
to be equal to x over a large band of wave numbers. But this alone does not guarantee that the scheme
has low numerical dispersion when used for wave propagation computation. To ensure low numerical
dispersion, the scheme must be such that d¯/d is nearly equal to 1 over a wide range of wave numbers.

4. Numerical dissipation

Discretization of a partial differential equation into a finite difference equation, generally, leads to
numerical dissipation in addition to numerical dispersion. Let us again consider the solution of the
convective wave equation (10) using a large stencil finite difference scheme. Suppose we approximate the
spatial derivative by an unsymmetric finite difference quotient and solve time exactly. The exact solution
of the semi-discretized problem is again given by (19). The integrand has a pole at  = c¯() in the
-plane. By means of the Residue Theorem, the -integral may be evaluated to give
 ∞
u(x, t) = ˆ () ei(x−c¯()t) d.
 (25)
−∞

Now whether the numerical solution is damped or not depends critically on whether a central difference
stencil or an unsymmetric difference stencil is used. If a central difference stencil is used, then ¯ () is a
real function for real . In this case, (25) represents a dispersive wave packet without being damped in
time. If an unsymmetric stencil is used, then ¯ () is complex for real  (see (5)). In this case, the solution
is damped in time provided c Im(¯) is negative for all . The time rate of damping for wave with wave
number  is given by c Im[¯()]. For many popular upwind numerical schemes c Im(¯) is negative. Such
schemes have built-in numerical damping. The damping rate can be calculated precisely once the stencil
size and stencil coefficients are known.
It is important to point out that it is possible c and Im(¯) have the same sign. In this case, the numerical
solution will grow exponentially in time leading to numerical instability. Therefore, a necessary condition
for numerical stability is
c Im[¯()] < 0, − x . (26)
This is the upwinding requirement. It is easy to show numerically that only upwind unsymmetric
stencils could satisfy condition (26). The upwinding requirement is extremely important when solving
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 603

the full Euler equations. The Euler equations support several modes of waves (acoustic, vorticity and
entropy waves). The upwinding condition must be satisfied by each wave mode if the numerical solution
is to remain stable.
Numerical damping can also be caused by temporal discretization. A more detailed discussion may be
found in Tam and Webb (1993).

5. Artificial selective damping

Numerical waves with wave number  for which ¯ () is not nearly equal to  (see Fig. 2) will not
propagate with the correct wave speed. For waves on the right side of the maximum of the ¯ versus 
curve of Fig. 2, the wave speed is negative (because d¯/d is negative) or opposite to the correct wave
propagation direction. These are the spurious waves of the computation scheme (Tam et al., 1993). They
must be removed from the computation if a high quality numerical solution is desired. A way to do this is
to apply artificial selective damping to the discretized equations. Suppose we add a damping term, D(x),
to the right side of the momentum equation of the linearized Euler equations in one dimension
ju 1 jp
+ = D(x). (27)
jt 0 j x
Let us discretize the spatial derivative using the 7-point optimized stencil as

du 1 
3
a 
3
+ aj p+j = − dj u+j . (28)
dt 0 x (x)2
j =−3 j =−3

In (28) it is assumed that D , the damping term, is proportional to the values of u (the variable with time
derivative) within the stencil. dj ’s are the damping stencil coefficients and a is the artificial kinematic
viscosity. a /(x)2 has the dimension of (t)−1 , so that dj ’s are pure numbers. Now we wish to choose
dj ’s so that the artificial selective damping would be concentrated mainly on the high wave numbers or
short waves.
The Fourier transform of the generalized continuous form of (28) is,
dũ a
+ ··· = − D̃(x)ũ, (29)
dt (x)2
where

3
D̃(x) = dj eij x . (30)
j =−3
On ignoring the terms not shown in (29), the solution is

ũ ∼ e−( a /(x)
2 )D̃(x)t
. (31)
Since D̃(x) depends on the wave number, the damping will vary with wave number. For our need,
we like D̃ to be zero or small for small x but large for large x. This can be done by choosing dj
appropriately. Tam et al. (1993) suggested that D̃ should be a positive even function of x. They used a
604 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

Fig. 7. D̃(x) versus x (damping curve).  = 0.3, 7-point stencil.

Gaussian function as a template to determine dj ’s. Their analysis gives the following values ( = 0.3):
d0 = 0.3276986608,

d1 = d−1 = −0.235718815,

d2 = d−2 = 0.0861506696,

d3 = d−3 = −0.0142811847.
A plot of D̃ versus x for this choice is shown in Fig. 7. Fig. 7 indicates that the grid-to-grid oscillations
with x =  or wavelength equal to 2 mesh spacings is most heavily damped. On the other hand, x = 0
and the low wave number waves are hardly damped at all. Experience has shown that the adoption of
artificial selective damping in a computation scheme can effectively eliminate all spurious short waves
in the numerical solution.
Damping stencil given by Fig. 7 is especially effective when the solution has discontinuities. It is very
useful for shock capturing (Tam and Shen, 1993). For general background damping, we prefer a damping
curve that has negligible damping up to x = 0.8. For this purpose, Tam et al. (1993) recommended the
following 7-point damping stencil coefficients ( = 0.2):
d0 = 0.2873928425,
d1 = d−1 = −0.2261469518,
d2 = d−2 = 0.1063035788,
d3 = d−3 = −0.0238530482.
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 605

Near a computation boundary where a 7-point stencil does not fit, one may use a 5-point and a 3-point
damping stencil. The coefficients of an optimized 5-point stencil and a 3- point stencil are given below:
5-point damping stencil
d0 = 0.375,

d1 = d−1 = −0.25,

d2 = d−2 = 0.0625,
7-point damping stencil
d0 = 0.5,

d1 = d−1 = −0.25.
To show the effectiveness of artificial selective damping, let us consider the solution of the convective
wave equation (10) again. Suppose the initial condition is a boxcar function
(x) = [H (x + 50) − H (x − 50)],
where H (x) is the unit step function. This initial condition has two sharp discontinuities. They are the
sources of spurious short waves.
Fig. 8(a) shows the computed solution using the 7-point stencil DRP scheme. As can be seen, the
solution is totally contaminated by spurious short trailing waves. To eliminate these spurious waves,
we may add artificial selective damping to the discretized form of (10). Fig. 8(b) shows the computed
solution after the pulse has propagated for 250 mesh spacings. In this calculation, the inverse artificial
mesh Reynolds number, R = cx/ a , is set equal to 1.0. The damping curve with  = 0.3 is used in
the computation. It is clear from Fig. 8(b) that the addition of artificial selective damping completely
eliminates the spurious spatial oscillations. The resulting solution is in good agreement with the exact
boxcar solution. Artificial selective damping, invariably, smooths out a discontinuity over several mesh
spacings. This is generally true for all shock capturing schemes. The only way to maintain a sharp
discontinuity is to do a shock/discontinuity fitting. Such an approach is difficult to carry out in two or
three-dimensional problems. It is seldom used.

6. Temporal discretization

There are two types of explicit time-marching schemes. They are:


(1) Single-step scheme, e.g., Runge–Kutta method.
(2) Multi-step scheme, e.g., Adams–Bashforth method.
Both types of methods are discussed in standard textbooks. Here, we will only discuss the optimized
multi-step method. One important advantage of multi-step method is that it can be used in a multi-size-
mesh multi-time-step algorithm (Tam and Kurbatshkii, 2003). Such an algorithm is very efficient and
nearly optimal in computation time.
Suppose u(t) is the unknown vector. The time axis is divided into a uniform grid with time step t.
We will assume that the values of u and du/dt are known at time level n, n − 1, n − 2, n − 3. (Note: In
606 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

Fig. 8. Computed solution of the convective wave equation with box car initial condition R −1 = 1.0,  = 0.3 damping curve
used in (b).

CAA, du/dt is given by the governing equations of motion.) To advance to the next time level, Tam and
Webb (1993) used the following 4-level finite difference approximation
 3  (n−j )
ju
u (n+1)
= u + t
(n)
bj , (32)
jt
j =0

where t is the time step. bj ’s are the coefficients of the scheme. The last term of (32) may be regarded
as the weighted average of the time derivatives of the last 4 mesh points. Finite difference equation (32)
may now be generalized to one with a continuous variable following the same steps that lead to (2). We
may apply Laplace transform to the generalized form of (32). On recalling that the Laplace transform
of du /dt is −iū , it is straightforward to obtain, assuming zero initial condition, (see Tam and Webb,
1993 for details and for cases with nonzero initial conditions)
−iū
− i¯ ū , (33)
where ¯ is given by
(e−it − 1)
¯ t = i 3
 . (34)
ij t
j =0 bj e
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 607

Since  on the left side of (34) is the angular frequency of the Laplace transform of d/dt, ¯ on the right
side may be regarded as the angular frequency of the 4-level time matching scheme (32). It is easy to
show numerically that  ¯ t is a good approximation of t for small t. For the DRP scheme, the
coefficients bj of (32) are chosen to minimize the difference between  ¯ t and t over a band of t.
Tam and Webb (1993) suggested the following numerical values:

b0 = 2.3025580888, b1 = −2.4910075998,

b2 = 1.5743409332, b3 = −0.3858914222.

They found that ¯ ≈  to a high degree of accuracy for t  0.19.

7. Radiation and outflow boundary conditions

Many interesting acoustic problems are exterior problems. To simulate this class of problems, it is
necessary to impose radiation and outflow boundary conditions at the boundaries of the finite computation
domain. To ensure that the computed solutions are of high quality, these boundary conditions must be
sufficiently transparent to the outgoing disturbances so that they exit the computation domain without
significant reflections. An alternative is to use absorbing boundary conditions such as perfectly matched
layer (PML) (see Hu, 1996, 2001; Tam et al., 1998). A high quality absorbing boundary condition absorbs
totally all outgoing waves without reflection. As is well known, the linearized Euler equations can support
three types of waves. Thus, in general, the outgoing disturbances would contain a combination of acoustic,
entropy and vorticity waves each having a distinct wave propagation characteristic. Here a set of radiation
and outflow boundary conditions compatible with the optimized high resolution scheme is developed
starting with the asymptotic solution. A review on numerical boundary condition can be found in Tam
(1998).
Consider two-dimensional exterior problems involving a uniform flow of velocity u0 in the x-direction
and sound speed a0 past some arbitrary acoustic, vorticity and entropy sources as shown in Fig. 9. It will
be assumed that the boundaries of the computation domain are quite far from the sources. At boundaries
where there are only outgoing acoustic waves the asymptotic solution (r → ∞), in polar coordinates
(r, ), is given by the following formula (see Tam and Webb, 1993),
⎡ 1 ⎤
⎡ ⎤ ⎡ ⎤ ⎢ a02 ⎥
a ⎢ ⎥
⎢ û() ⎥
⎢ u ⎥ ⎢ ua ⎥ F (r/V () − t, ) ⎢ ⎥
⎣ ⎦≡⎣ ⎦= ⎢ 0 a0 ⎥ + O(r −3/2 ), (35)
v va r 1/2 ⎢ ⎥
⎢ v̂() ⎥
p pa ⎣ ⎦
0 a0
1

where V () = a0 [M cos  + (1 − M 2 sin2 )1/2 ] and M is the mean flow Mach number. The subscript
‘a’ in ( a , ua , a , pa ) above indicates that the disturbances are associated with the acoustic waves alone.
By taking the time t and r derivatives of (35) it is straightforward to find that for arbitrary function F , the
608 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

Fig. 9. Acoustic, vorticity and entropy wave sources in a uniform flow.

acoustic disturbances satisfy


⎡ ⎤
 
1 j j 1 ⎢u⎥ −5/2
+ + ⎣ ⎦ = 0 + O(r ). (36)
V ( ) jt jr 2r v
p
Eq. (36) provides a set of far field radiation boundary conditions.
At the outflow region, the outgoing disturbances, in general, consist of a combination of acoustic,
entropy and vorticity waves. The asymptotic solutions for the density, velocity and pressure fluctuations
are given by (Tam and Webb, 1993),
⎡ (x − u t, y) + ⎤
⎡ ⎤ 0 a
⎢ j (x − u t, y) + u ⎥
⎢u⎥ ⎢ jy
0 a ⎥
⎣ ⎦=⎢ ⎢ j
⎥ + ··· ,
⎥ (37)
v ⎣ − (x − u0 t, y) + va ⎦
p jx
pa
where the explicit form of ( a , ua , a , pa ) may be found in (35). The functions ,  and F are entirely
arbitrary. These functions can be eliminated by a combination of derivatives. In this way, the following
outflow boundary conditions are derived.
 
j j 1 jp jp
+ u0 = 2 + u0 ,
jt jx a0 jt jx

ju ju 1 jp
+ u0 =− ,
jt jx 0 jx
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 609

jv jv 1 jp
+ u0 =− ,
jt jx 0 jy
1 jp jp jp p
+ cos  + sin  + = 0. (38)
V () jt jx jy 2r
For three-dimensional problems such as jet flows, where the mean flow profiles at the outflow boundary
is not known, the outflow boundary conditions must be capable of capturing the mean flow as well as
allowing outgoing disturbances to exit the computation domain without reflection. For these purposes,
one may nonlinearize the convective terms of the above set of equations to form a set of outflow boundary
conditions. In three-dimensional cylindrical coordinates (r, , x) , they are
 
j j j w j 1 jp jp jp w jp
+u +v + = +u +v + , (39)
jt jx jr r j a 2 jt jx jr r j
ju ju ju w ju 1 jp
+u +v + =− , (40)
jt jx jr r j jx

jv jv jv w jv w2 1 jp
+u +v + − =− , (41)
jt jx jr r j r jr
jw jw jw w jw vw 1 1 jw
+u +v + + =− , (42)
jt jx jr r j r r j
1 jp jp p − p∞
+ + = 0, (43)
V () jt jR R
where (R, , ) in (43) are the spherical polar coordinates with the polar axis coinciding with the centerline
of the jet and the x-axis.

8. Applications

The goal of CAA is not just to develop new methods to solve or simulate aeroacoustics problems. An
important activity of CAA is in applications. CAA should not be viewed as just methods for providing
numerical results and predictions alone. CAA simulations offer a complete set of space-time data of an
aeroacoustic phenomenon. One could make use of the data to investigate the noise generation mechanisms
and the physics involved in much the same way as physical experiments.
One intrinsic difference between aeroacoustics and fluid dynamics problems is that the former are
generally multi-scale problems. Here we will consider an application of CAA to illustrate the multi-scale
nature of aeroacoustics problems.
Presently, all commercial jet engines are, invariably, fitted with acoustic liners for fan noise suppression.
A jet engine acoustic liner usually consists of a face sheet full of small holes (Motsinger and Kraft, 1991).
The holes are usually arranged in a regular pattern. Underneath the face sheet are cavities, sometimes
called resonators. When sound waves impinge on the surface of a liner, fluid is forced into the cavities
through the holes during the high pressure half of a cycle. During the low pressure half of a cycle,
fluid flows back out from the cavities. Viscous damping of the oscillating flow in and out of the holes
610 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

Fig. 10. (a) Vortex shedding at a 90◦ slit. (b) Vortex shedding and the development of thin shear layers and subsequent rolling
up into vortices due to Kelvin–Helmholtz instability at a 45◦ beveled slit.

connecting the outside to the cavities has been regarded as the mechanism responsible for the dissipation
of the incident sound. Because of the smallness of the holes, typically 1 mm or less in diameter, there has
not been experimental observation of the actual flow and acoustic fields around the holes.
In the past, some investigators suggested that wall friction was responsible for acoustic dissipation.
Based on experiments carried out using larger size holes (larger Reynolds number), others have suggested
that the oscillatory flow through the holes led to the formation of oscillatory turbulent jets. Acoustic
dissipation is the result of the conversion of acoustic energy to fluid turbulence, which is subsequently
dissipated by molecular viscosity.
Tam and Kurbatshkii (2000) and Tam et al. (2001a,b, 2005, 2006) noted that the small size of the
holes of an acoustic liner is of no hindrance to numerical simulation. They performed DNS of a slit
resonator (a single slit and a single cavity) under acoustic excitation to investigate the mechanisms
by which a perforated acoustic liner dissipates sound. Tam and Kurbatshkii (2000) found from their
numerical simulations that at low incident sound pressure level, a jet-like oscillatory shear layer was
formed near each side wall of the slit. The viscous dissipation associated with these jet-like shear layers
was primarily responsible for acoustic energy dissipation. At high incident sound pressure level, they
observed vortex shedding at the corners of the slit opening connecting the resonator to the outside (see
Fig. 10a). There was a large increase in dissipation whenever there was vortex shedding. The dominant
dissipation mechanism was the conversion of acoustic energy into the rotational kinetic energy of the
shed vortices. These vortices were subsequently dissipated into heat by molecular viscosity. The vortex
shedding mechanism was new and the finding represented a genuine CAA contribution. To ensure that
the numerical results were valid, physical experiments were performed in parallel in Tam et al. (2005,
2006). In the experiments, gross quantities such as impedance and reflection coefficients, but not detailed
flow field, were measured. The computed reflection coefficients and impedance over a wide range of
frequencies, slit widths and slit geometry (see Fig. 10b for beveled slit) were found to be in excellent
agreement with the companion experimental measurement offering confidence in the accuracy of the
flow and acoustic field including vortex shedding provided by the numerical simulation but not directly
measured or observed in the experiments.
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 611

There are two dominant dynamical effects in the acoustic liner problem. Near the opening of the slit
or hole, viscous forces are important. Away from the opening, compressibility effect is most important.
Adjacent to the walls of the opening, the oscillatory fluid motion induces the formation of a viscous
Stokes layer (see White, 1991, Chapter 3). The wavelength, , of an oscillatory Stokes layer is
 1/2

=2 , (44)
f
where is the kinematic viscosity and f is the frequency. Suppose the 7-point stencil DRP scheme is
used for computation. Since this scheme can resolve waves using 7 or 8 mesh points per wavelength,
the mesh size required at the slit is (x)Stokes = /8 = 0.8( /f )1/2 . Outside the slit opening region, the
fluid motion is dominated by compressibility effect. The acoustic wavelength is a0 /f where a0 is the
speed of sound. To meet resolution requirement, the mesh spacing must, therefore, be equal to or less
than (x)acoustic = a0 /(8f ). The ratio of mesh size is
(x)acoustic 0.277a0
= (45)
(x)Stokes ( f )1/2
under standard conditions, this ratio is equal to 640 for 3 kHz sound. It would make no sense to use a
uniform size mesh for DNS. In the work of Tam et al. (2005, 2006), the multi-size-mesh multi-time-step
DRP scheme of Tam and Kurbatshkii (2003) was used. The computation domain was partitioned into
eight subdomains inside and outside the resonator. Thus the largest size mesh was 27 times larger than
the smallest size mesh. This also applied to the time steps. Effectively, more than half of the computation
was performed in the three subdomains with the finest meshes.
There are interesting CAA contributions in other areas of applications. In jet mixing noise simulation,
significant advances were made by Bogey et al. (2002), Freund et al. (2000), Freund (2001) and Morris
et al. (1999). In jet screech study, the phenomenon was reproduced in the numerical simulations of
Shen and Tam (1998, 2000, 2002). Manning and Lele (1998, 2000) and Loh et al. (2001) independently
investigated jet screech with slight different objectives. In airframe noise, Tam and Pastouchenko (2001)
investigated the mechanism by which gap tones were generated by DNS. Numerical simulation of cavity
tone generation has recently become quite successful. Readers interested in these studies may consult the
works of Rowley et al. (2002), Shieh and Morris (2001), Shih et al. (1994), Zhang and Edwards (1990),
and Zhang (1995).

9. Summary and concluding remarks

In this paper, the needs and unique challenges of CAA are discussed. They are different from those that
motivated the development of CFD. There is no question that CFD has been very successful in solving
aerodynamic and standard fluid dynamic problems. Yet CFD methods are not designed for computing
aeroacoustic problems. As a result, an independent development of CAA was necessary during the last
decade.
Many of the original computational challenges confronting CAA have now been adequately resolved.
They are discussed in this paper. They include the development of high resolution spatial discretization
schemes using no more than 4–8 mesh points per wavelength. A method to develop optimized temporal
discretization schemes has also been discussed. Two of the most important errors in wave propagation
612 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

computations are numerical dispersion and dissipation errors. In this paper, the origins of numerical
dispersion and dissipation are rigorously analyzed in wave number space through Fourier-Laplace trans-
forms. High resolution numerical schemes, invariably, support spurious short waves which, when excited,
would contaminate the computed solution. One way to eliminate these spurious waves is to use artificial
selective damping. A method to determine artificial selective damping stencil coefficients is discussed in
details. A typical CAA algorithm consists of a high resolution time marching scheme, a well designed
computation grid, a properly chosen set of artificial selective damping stencils or filters and a set of
high quality radiation, inflow and outflow boundary conditions. Here the construction of high quality
asymptotic radiation and outflow boundary conditions is discussed at some length.
As CAA matures and the scope of applications broadens, new pressing challenges emerge. Among these
are ways to treat complex geometry with consistent high resolution in wave number space throughout
the entire computation domain. In CFD, the most popular methods for treating complex geometry are
the unstructured grids and the overset grids. Unfortunately, unstructured grids are most effective and
convenient when used in conjunction with finite volume method. Finite volume method is essentially a
low-order scheme (first- or second-order). It is not suitable for solving aeroacoustic problems. On the
other hand, significant progress has recently been made in the development of overset grid method for
CAA. Delfs (2001), Yin and Delfs (2001) were the first to explore this possibility. The work of Sherer and
Visbal (2003) represents a significant advancement of the method. Tam and Hu (2004) recognized that
for overset grids to succeed, it was imperative that the data transfer process must maintain the same high
resolution requirement as the basic time marching computation method. Accordingly they developed an
optimized high resolution interpolation method specifically for overset grids applications.
There is also a need to develop outflow boundary conditions that can capture the mean flow accurately
as well as being transparent to outgoing disturbances. Many CAA outflow boundary conditions are
developed assuming that the mean flow is already known. This is true for both asymptotic outflow
boundary conditions or absorbing boundary conditions such as perfectly matched layer (PML). In most
practical problems the mean outflow is not known a priori. Computing the unknown mean outflow is an
important goal in many of these problems.
In aeroacoustics, turbulence is the principal source of broadband noise. Therefore, research and de-
velopment of turbulence modeling and turbulence simulation are an integral part of CAA. It is known
that direct numerical simulation (DNS) of high Reynolds number turbulent flows requires exceedingly
large number of mesh points and large CPU time. On account of such requirements, DNS is, presently,
not considered feasible for solving practical CAA problems. Recently, attention has turned to large eddy
simulation (LES). However, owing to the three-dimensional nature of turbulence, realistically, LES com-
putation can be carried out only in relatively small computation domains in most instances. Simple
estimates of mesh and computer time requirements would convince most that it would be sometime in
the future, when much larger and faster computers become available, before LES would become a design
tool for CAA.
In CFD, calculating the mean velocity profile and other mean quantities of turbulent flows is important
for engineering applications. For mean flow calculation, one may use a two-equation turbulence model
(e.g., the k . or the k . model) or, if desired, a more advanced model. The pertinent question is whether
models of a similar level of sophistication could be developed for noise calculation. Recent efforts of this
kind have made significant progress in jet noise prediction. Tam and Auriault (1999) investigated how
fine scale turbulence in a jet generates noise. They observed an analogy between pressure created by the
random motion of blobs of fine scale turbulence and pressure created by random motion of gas molecules
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 613


in gas kinetic theory. They proposed that since q = 23 k, where k is the turbulence kinetic energy per
unit mass, is the turbulence pressure according to the analogy, the dominant noise source of fine scale
turbulence is the time rate of change of q in the fluid frame of motion. Starting from such a proposition, Tam
and Auriault developed a fine scale turbulence noise theory for jets. The turbulence information needed
by the theory including turbulence intensity, length and time scales, were provided by the k . turbulence
model. The noise spectra predicted by the Tam and Auriault theory were found to agree well with measured
data over a wide range of jet Mach number for axisymmetric jets as well as nonaxisymmetric jets (Tam
and Pastouchenko, 2002) and jets in simulated forward flight (Tam et al., 2001b). Recently, the theory
was found to be capable of predicting the noise source strength distribution along the axial distance of
jets. The predicted distributions (Tam et al., 2003b) both for selected Strouhal number and noise intensity
(integrated over all frequencies) were in good agreement with the measurements of Laufer et al. (1976)
and Schlinker (1975). Based on the success of the effort of Tam and Auriault it seems reasonable to
suggest that, perhaps, CAA turbulence modeling effort should focus on how turbulence generates noise.
This is not a subject of high priority interest of the turbulence research community.
The Tam and Auriault theory is designed for predicting noise from fine scale turbulence in free shear
flows. However, for broadband fan noise and airframe noise the interaction of turbulence and solid surfaces
is a major source of broadband noise. There is a critical need for such a theory. Any fresh idea that might
pave a way to the development of such a theory would be a significant contribution.

Acknowledgements

This work was supported by NASA Grant NAG1-2145, NASA Contracts NAS1-02045, NAS3-02125
and the Aeroacoustics Research Consortium.

References

Ablowitz, M.J., Fokas, A.S., 1997. Complex Variables: Introduction and Applications. Cambridge University Press, Cambridge.
Bogey, C., Bailly, C., Juve, D., 2002. Noise investigation of a high subsonic, moderate Reynolds number jet using a compressible
LES. AIAA Paper 2000-2009.
Delfs, J.W., 2001. An overlapped grid technique for high resolution CAA schemes for complex geometries. AIAA Paper
2001-2199.
Freund, J.B., 2001. Noise sources in a low-Reynolds number turbulent jet at Mach 0.9. J. Fluid Mech. 438, 277–305.
Freund, J.B., Lele, S.K., Moin, P., 2000. Numerical simulation of a Mach 1.92 turbulent jet and its sound field. AIAA J. 38,
2023–2031.
Hsi, M.Y., Perie, F., 1977. Computational aeroacoustics for prediction of acoustic scattering. In: Tam, C.K.W., Hardin,
J.C. (Eds.), Proceedings Second Computational Aeroacoustics Workshop on Benchmark Problems, NASA CP 3352.
pp. 111–117.
Hu, F.Q., 1996. On absorbing boundary conditions of linearized Euler equations by a perfectly matched layer. J. Comput. Phys.
129, 201–219.
Hu, F.Q., 2001. A stable perfectly matched layer for linearized Euler equations in unsplit physical variables. J. Comput. Phys.
173, 453–480.
Laufer, J., Schlinker, R.H., Kaplan, R.E., 1976. Experiments on supersonic jet noise. AIAA J. 14, 489–497.
Lele, S.K., 1992. Compact finite difference schemes with spectral-like resolution. J. Comput. Phys. 103, 16–42.
Lele, S.K., 1997. Computational aeroacoustics: a review. AIAA Paper 97-0018.
Li, Y., 1997. Wave number-extended high-order upwind-biased finite difference schemes for convective scalar transport.
J. Comput. Phys. 133, 235–255.
614 C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615

Lockard, D.P., Brentner, K.S., Atkins, H.L., 1995. High-accuracy algorithms for computational aeroacoustics. AIAA J. 33,
246–251.
Loh, C.Y., Hultgren, L.S., Jorgenson, P.C.E., 2001. Near field screech noise computation for an underexpanded supersonic jet
by the CE/SE method. AIAA Paper 2001–2252.
Manning, T.A., Lele, S.K., 1998. Numerical simulations of shock vortex interactions in supersonic jet screech. AIAA Paper
1998-0282.
Manning, T.A., Lele, S.K., 2000. A numerical investigation of sound generation in supersonic jet screech. AIAA Paper 2000-2081.
Morris, P.J., Long, L.N., Scheidegger, T., 1999. Parallel computations of high speed jet noise. AIAA Paper 99-1873.
Motsinger, R.E., Kraft, R.E., 1991. Design and performance of duct acoustic treatment. In: Hubbard, H.H. (Ed.), Aeroacoustics
of Flight Vehicles: Theory and Practice, NASA RP-1258, vol. 2. pp. 165–206.
Rowley, C.R., Colonius, T., Basu, A.J., 2002. On self-sustained oscillations in two-dimensional compressible flow over
rectangular cavities. J. Fluid Mech. 455, 315–346.
Schlinker, R.W., 1975. Supersonic jet noise experiments. Ph.D. Thesis, Department of Aerospace Engineering, University of
Southern California.
Shen, H., Tam, C.K.W., 1998. Numerical simulation of the generation of axisymmetric mode jet screech tones. AIAA J. 36,
1801–1807.
Shen, H., Tam, C.K.W., 2000. Effects of jet temperature and nozzle-lip thickness on screech tones. AIAA J. 38, 762–767.
Shen, H., Tam, C.K.W., 2002. Three-dimensional numerical simulation of the jet screech phenomenon. AIAA J. 40, 33–41.
Sherer, S.E., Visbal, M.R., 2003. Computational study of acoustic scattering from multiple bodies using a high-order overset
grid approach. AIAA Paper 2003-3203.
Shieh, C.M., Morris, P.J., 2001. Comparison of two- and three-dimensional turbulent cavity flows. AIAA Paper 2001-0511.
Shih, S.H., Hamed, A., Yenan, J.J., 1994. Unsteady supersonic cavity flow simulations using coupled k . and Navier–Stokes
equations. AIAA J. 32, 2015–2021.
Tam, C.K.W., 1995. Computation aeroacoustics: issues and methods. AIAA J. 33, 1788–1796.
Tam, C.K.W., 1998. Advances in numerical boundary conditions for computational aeroacoustics. J. Comput. Acoust. 6,
377–402.
Tam, C.K.W., Auriault, L., 1999. Jet mixing noise from fine scale turbulence. AIAA J. 37, 145–153.
Tam, C.K.W., Hu, F.Q., 2004. High resolution overset grid method for computational aeroacoustics, AIAA Paper 2004-2812.
Tam, C.K.W., Kurbatshkii, K.A., 2000. Microfluid dynamics and acoustics of resonant liners. AIAA J. 38, 1331–1339.
Tam, C.K.W., Kurbatshkii, K.A., 2003. Multi-size-mesh multi-time-step dispersion-relation-preserving scheme for multiple-
scales aeroacoustics problems. Int. J. Comput. Fluid Dyn. 17, 119–132.
Tam, C.K.W., Pastouchenko, N.N., 2001. Gap tones. AIAA J. 39, 1442–1448.
Tam, C.K.W., Pastouchenko, N.N., 2002. Noise from fine scale turbulence of non-axisymmetric jets. AIAA J. 40, 456–464.
Tam, C.K.W., Shen, H., 1993. Direct computation of nonlinear acoustic pulses using high-order finite difference schemes. AIAA
Paper 93-4325.
Tam, C.K.W., Webb, J.C., 1993. Dispersion-relation-preserving finite difference schemes for computational acoustics. J. Comput.
Phys. 107, 262–281.
Tam, C.K.W., Webb, J.C., Dong, Z., 1993. A study of the short wave components in computational acoustics. J. Comput. Acoust.
1, 1–30.
Tam, C.K.W., Auriault, L., Camballi, F., 1998. Perfectly matched layers as an absorbing boundary condition for the linearized
Euler equation in open and ducted domains. J. Comput. Phys. 144, 213–234.
Tam, C.K.W., Kurbatskii, K.A., Ahuja, K.K., Gaeta Jr., R.J., 2001a. A numerical and experimental investigation of the dissipation
mechanisms of resonant acoustic liners. J. Sound Vib. 245, 545–557.
Tam, C.K.W., Pastouchenko, N.N., Auriault, L., 2001b. The effects of forward flight on jet mixing noise from fine scale turbulence.
AIAA J. 39, 1261–1269.
Tam, C.K.W., Ju, H., Jones, M.G., Watson, W.R., Parrot, T.L., 2005. A computational and experimental study of slit resonators.
J. Sound Vib. 284, 947–984.
Tam, C.K.W., Pastouchenko, N.N., Schlinker, R.H., 2006. On the two sources of supersonic jet noise. J. Sound Vib. 291, 192–201.
Trefethen, L.N., 1982. Group velocity in finite difference schemes. Siam Rev. 24, 113–136.
Vichnevetsky, R., Bowles, J.B., 1982. Fourier Analysis of Numerical Approximations of Hyperbolic Equations. SIAM,
Philadelphia.
Wells, V.L., Renaut, R.A., 1997. Computing aerodynamically generated noise. Ann. Rev. Fluid Mech. 29, 161–199.
C.K.W. Tam / Fluid Dynamics Research 38 (2006) 591 – 615 615

White, F.M., 1991. Viscous Fluid Flow. McGraw-Hill, New York.


Whitman, G.B., 1974. Linear and Nonlinear Waves. Wiley-Interscience, New York.
Yin, J., Delfs, J.W., 2001. Sound generation from gust-airfoil interaction using CAA-chimera method. AIAA Paper 2001-2136.
Zhang, X., 1995. Compressible cavity flow oscillation due to shear layer instabilities and pressure feedback. AIAA J. 33,
1404–1411.
Zhang, X., Edwards, J.A., 1990. An investigation of supersonic oscillatory cavity flows driven by a thick shear layer. Aeronaut.
J. 94, 355–364.
Zhuang, M., Chen, R., 1998. Optimized upwind dispersion-relation-preserving finite difference scheme for computational
aeroacoustics. AIAA J. 36, 2146–2148.
Zhuang, M., Chen, R., 2002. Applications of high-order optimized upwind schemes for computational aeroacoustics. AIAA
J. 40, 443–449.
Zingg, D.W., 2000. Comparison of high-accuracy finite difference methods for linear wave propagation. SIAM J. Sci. Comput.
22, 476–502.

You might also like