Sie sind auf Seite 1von 14

47th AIAA Aerospace Sciences Meeting Including The New Horizons Forum and Aerospace Exposition

5 - 8 January 2009, Orlando, Florida

AIAA 2009-10

47th Aerospace Sciences Meeting and Exhibit


58 January 2009, Reno, NV

Characteristic boundary conditions for


non-orthogonal, moving meshes
Daniel J. Bodony
Department of Aerospace Engineering
University of Illinois at Urbana-Champaign
Urbana, IL 61801
Boundary conditions to the compressible Navier-Stokes equations are developed for the case of deformable,
generalized coordinates. The general theory is based on a idea of Halpern [SIAM J. Math. Analy., Vol. 22(5),
pp. 12561283, 1991] which, in the inviscid case, reduces to standard characteristic treatment and thus logically
extends the work of Thompson [J. Comput. Phys., vol. 68, pp. 124, 1987], Poinsot & Lele [ibid, vol. 101,
pp. 104129, 1992], and Kim & Lee [AIAA J., vol. 42(1), pp. 4755, 2004]. The issue of well-posedness is
considered. The developed boundary conditions are applicable to fluid problems with moving boundaries in
inviscid and viscous fluids. Several verification problems are presented to demonstrate accuracy.

I. Introduction
The simulation of unsteady fluid flows has had a profound influence on the development of boundary conditions
that can accurately represent the required boundary data. One of the requirements that all boundary condition implementations should satisfy is the correct specification of the physical boundary conditions as determined by analysis,1, 2
with the remaining numerical boundary conditions specified accordingly.
For inviscid flows the boundary conditions may be constructed based on characteristics, such as considered by
Thompson3, 4 and by Poinsot & Lele.5 For flows with non-zero viscosity the notion of characteristics fails as the governing equations are no longer hyperbolic, often being called incompletely parabolic. As such characteristic boundary
conditions are not strictly applicable but, in practice, have been applied with success in a number of cases. In these
instances the issue of well-posedness is not certain and the failure of a simulation to remain stable, in the sense of
Kreiss,6 suggests that it may be the inconsistent boundary condition formulation causing the instability.
More complete boundary conditions for the compressible viscous governing equations have been developed by
Hesthaven & Gottlieb,7 Svard et al.,8 and Svard & Nordstrom,9 among others, based on a penalization technique
(called Simultaneous Approximation Term, or SAT) combined with very specific internal finite difference operators.
These formulations, which are based on local data, have been shown to be well-posed2 but introduce at least one
artificial time scale into the problem through the penalty parameter. These schemes have also been shown to preserve
the formal accuracy of the overall numerical method. These conditions are imposed weakly in the equations as


q
= + (A+ + B)q g
t

where = I + V is usually a scalar with specific bounds to ensure stability. A+ is the incoming part of the
hyperbolic characteristics, B is a matrix associated with the viscous influence on the boundary condition through the
small parameter , and g are the boundary data to be imposed. The reader is referred to Refs. 8, 9 for the details.
The introduction of SAT-based boundary conditions into an existing code requires the implementation of new
finite difference schemes following the Summation-by-parts (SBP) property.10 Together, the SBP-SAT formulation
can be shown to yield provably stable numerical discretizations for the compressible Navier-Stokes equations on
uniform Cartesian meshes. For non-uniform meshes only the diagonal SBP schemes up to and including 3rd order

AIAA member. bodony@illinois.edu


c 2009 by D. J. Bodony. Published by the American Institute of Aeronautics and Astronautics, Inc. with permission.
Copyright

1 of 14
InstituteofofAeronautics
Aeronauticsand
andAstronautics,
AstronauticsInc.,
Paper
Copyright 2009 by Daniel J. Bodony. Published by the American
American Institute
with2009-0010
permission.

admit an energy norm-based stability proof. When one is interested in a numerical method higher than 3rd order
or if one wishes to use alternative finite difference schemes there is no advantage to the SBP-SAT approach other
than through the existence of well-posed boundary conditions. As such one must search for other formulations. As
mentioned earlier the inviscid equations are well-posed when coupled with characteristic-based boundary conditions,
such as developed by Thompson.3, 4 The later development by Poinsot & Lele5 improved on Thompsons approach,
and included the viscous terms in an ad hoc manner. Further work by Kim & Lee12, 13 partially generalized the onedimensional approach by including the metric transformation for non-uniform meshes but also altered the manner
in which the wave amplitudes are calculated through the inclusion of the transverse terms, possibly violating well
posedness.
The approach taken by Poinsot & Lele and by Kim & Lee with respect to the viscous terms was arbitrary but
guided by the work of Engquist & Majda,14 Rudy & Strikwerda,15 and by Oliger & Sundstrom.1 Absent any proof of
well posedness the failure of these boundary conditions, in the sense of numerical instability, is never precisely known.
Thus the purpose of this work is to develop boundary conditions which are well posed and which may be cast in
the form consistent with Thompsons approach, in the sense that time-dependent equations are solved on the boundary
along with the interior scheme without the need for implementation of the SBP-SAT schemes. This is not a criticism
of SBP-SAT but an alternative to it. The continuous equations will be developed following the approach of Halpern.16
Since it may be shown that this approach is equivalent to the characteristic-based boundary conditions the discussion
will start with the more traditional development in the style of Thompson. Later the viscous boundary conditions will
be stated.

II. Derivation of the characteristic relations for an inviscid fluid


The conservation form of the governing equations of a compressible fluid, in Cartesian form, may be writtena
Q Fi
=S
+
t
xi

(1)

where Q = [ u E]T is the vector of mass, momentum, and energy density per unit volume. We take u to be the velocity vector in N-dimensions. The Fi are the fluxes in the ith direction and contain only the inviscid contributions FiI .
The Cartesian coordinates (x, t) can be mapped to another coordinate system (, ) via the time-dependent mappings
x = X(, )

with inverse

= (x, t)

(2)

where X 1 = and we only consider non-singular mappings such that X 1 exists and is well defined. Moreover we
take t = . The Jacobian of the transformation is defined as J = det(i /x j ) and is strictly positive.
Under these conditions and with simple application of the chain rule it can be shown17 that Eq. (1) tranforms to
i S
 Q  F
+
=
J
i
J

(3)

after using the identities

j
!

1
+
J
j

!
1 j
=0
J xi
!
1 j
= 0.
J t

for i = 1, . . . , N
(4)

If we define the weighted metric i = J 1 (/xi ) and contravariant velocity U = u j j + t , with similar expressions
for the remaining components, then the inviscid fluxes F iI are

a The

1I = uU + p x
F
vU + p

(E + p)U t p

and

2I = uV + p x
F
vV + p

(E + p)V t p

summation convention is used where repeated indices are summed from 1 to N.

2 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

(5)

in two dimensions and

uU + px

I
1 = vU + py ,
F

wU + p
z

(E + p)U t p

uV + p x

I
2 = vV + p y ,
F

wV + p

(E + p)V t p

and

in three dimensions.
Following Thompson we write Eq. (3) as

+ px
uW

I
3 = wV + py
F

wW
+ pz

t p
(E + p)W

!
Ij
F
1
Q
=SQ
+J
S

j
J

(6)

(7)

For the case with the coordinate = 1 is not in the plane of the boundary, but is not necessarily normal to the
boundary, we will move the boundary-parallel fluxes to the right-hand-side of Eq. (7) as
N
Ij
X
1I
F
F
Q

+J
= S J

j
j=2

(8)

into its Cartesian components,


and expand F
F I = j FIj + t Q

(9)

so that Eq. (8) becomes


(
)
)
(
F j
Q
Q

+ J j
+ t
= S J F j ( j ) + Q (t ) .

Upon introducing the matrix A

F j

A= J
+ t I

j Q

(10)

(11)

with I the (N + 2) (N + 2) identity matrix we can write Eq. (10) in the desired form
Q
Q
+A
=S

(12)

where

(
)

S = S J F j ( j ) + Q (t ) .
(13)

Equation (12) is the final form of the governing equations. It is nothing more than a rearrangment of the -fluxes to
the left-hand-side and the remaining terms to the right-hand-side.
From Eq. (12) we now split Q into its characteristic components. To do this we note that the submatrix
J j

FIj
Q

is diagonalized by the matrix P according to


J j

FIj
Q

= PP1

where the matrix P and its inverse may be found in Hirsch.18 The current matrix A contains the additional term t I.
FI
Since P diagonalizes J j j we may write
Q

=P

FIj

J j
+ J t I J t I P
Q

= P1 AP J t P1 IP
= P1 AP J t I.
3 of 14

American Institute of Aeronautics and Astronautics Paper 2009-0010

(14)

On defining U = J U = J(u j j + t ) to be the J-scaled contravariant velocity associated with we find that (from
Eq. (14)) we have
= P1 AP

(15)
where
= diag (U, U, U, U + c| x |, U c| x |)

(16)

and where x is the gradient of with respect to the Cartesian x coordinate.


we now diagonalize Eq. (12) as
Using P and
R R
+
= Sc

(17)

where the variations (in three dimensions) are




V,
p/(c) + U,
p/(c) U
R = p/c2 , W,

(18)

with U = x j u j the contravariant component of the velocity at the wall which does not include the term t . Here
x j = x j /| x | is the jth component of the unit vector in the direction of x . Likewise we write variations V =
= x w z u. Equation (17) is the characteristic form of the governing equations to which we
x v + y u and W
apply the appropriate boundary conditions. The elements of the vector
R
L=

(19)

can be associated (under suitable approximations) with the wave amplitudes of the incoming and outgoing components
of R.5 The expression for L in Eq. (19) reduces to the similar result found by Kim & Lee12 when the grid is fixed, i.e.,
when t 0. When, in addition, we take SC = 0 we recover the equivalent expression found by Poinsot & Lele when
expressed in curvilinear coordinates.

III. Boundary conditions for the viscous compressible Navier-Stokes


As mentioned in the introduction the use of characteristics for developing boundary conditions does not ensure
well posedness for the viscous equations, although in practice such approaches have been known to work reasonably
well for high Reynolds numbers. For example, consider the linear advection diffusion equation ut + au x u xx =
0, u(x, t = 0) = u0 for which the Cauchy problem is well posed. Split the x-axis into two parts, and consider the
problem to be solved on the left half domain = {(x, t)| R R+ }. Since the domain is semi-infinite one must
apply transparent boundary conditions on = {x = 0} to mimic the right-half real line. Loheac19 showed that one
could close the problem by solving ut + au x = 0 for x and get an error that is of order 2 as 0. Such a
result qualitatively explains the experience regarding characteristic-based boundary conditions for viscous problems.
For reacting flows Sutherland & Kennedy20 found that ignoring the viscous flux on the boundary, though stable, was
unsuitably inaccurate.
Thus one is left to search for more accurate, and well posed, boundary conditions. For viscous non-slip walls
Svard & Norstrom9 have shown that specifying the wall velocity and the wall temperature is well posed such that one
only needs to solve for the density (or other suitable thermodynamic variable) on the wall. The case when the normal
temperature gradient is specified is treated similarly.
When the boundary represents an open one, in the sense that waves may pass in or out as necessary, the situation
in more complex. Thus, following Halpern,16 consider perturbations w = [u T ]T about a uniform base state
[u T ]T . These perturbations satisfy
w
w
w
+ P( j,k)
(20)
= A( j)
t
x j
x j xk
where A( j) and P( j,k) are n n square matrices. The particular choice of w has been made so that the A and P may be
written

( j,k)

B( j) C ( j)
P
0
( j)
(
j,k)
, P
A = ( j)
=

( j)
A
D
0
0
( j)

where t A j is strictly hyperbolic, and A

( j)

is diagonal with p negative eigenvalues. The rank of P

4 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

( j,k)

is r.

In what follows a right boundary at x = 0 is considered on a Cartesian mesh. Other boundaries and mesh nonuniformity can be included without difficulty. The essential idea of the boundary conditions is written out as the following
steps
1. Fourier-Laplace transform in y = (x2 , x3 , . . . )T and in time, respectively. The Fourier variable is ; the Laplace
variable is s.
2. Use the transmission condition that
P

( j,1) w

is continuous over
x
w is continuous over

to develop boundary conditions in the Fourier-Laplace domain.


3. Expand the conditions in series of and , the transverse wavenumber
4. Invert the boundary conditions to local space-time following Engquist & Majda14, 21
The non-local, exact and well posed boundary conditions may be written
P

I
(1, j) w

x j

r+p
X

i=m+1

wk =

(1,1)

r+p
X

i
i Ni1
j jw j

(21)

j=1

r+p X
r+p
X

i
Ni1
j k w j ,

i=1 j=1

r+ p+1i n

(22)

where wI = [u T ]T , (i , i ) solve a generalized eigenvalue problem to be given below, and the columns of N are
constructed from portions of . Observe that the first row corresponds to boundary conditions specific to viscous
flows. Likewise it may be shown that the second row, when = 0, is identical to characteristic boundary conditions.
Construction of (i , i ) is as follows. Consider the generalized eigenvalue problem
(A(1) + 2 P(11) sI) = 0.

(23)

For the 1-D compressible Navier-Stokes equations this is quintic equation in . To each of the the n eigenvalues i and
eigenvectors i of A(1) are associated, for the case when = 0,
i (s, ) = i (s) + i s2 + O( 2 ),

i = s/i

and
i (s, ) = i + i s + O( 2 ),

i = i .

for i m where m is the number of negative eigenvalues i .


The viscous component satisfies
i (s, ) = i (s) + i s + O(),

i = i /

i (s, ) = i + i s + O( 2 ),

i = i .

and
where (, ) satisfy (P(1,1) + A(1) ) = 0. Observe that the eigenvalues i and eigenvectors i have been expanded
in terms of the small parameter and that = 0 has been assumed. The latter condition, which assumes wavefronts
parallel to boundary, can be relaxed to include obliquely traveling waves.
For the one-dimensional Navier-Stokes equations we find that

A(1)

= ( 1)T

a2 /(T )
u
0

a2 /

and

P(1,1) = diag(4/3, /Pr, 0)

5 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

from which is may be shown that

1 =

s
3( 1) + 4 2
+
s + O( 2 ),
a u
6(a + u)3

2 =

2
s
s + O( 2 ),
+
u u3

3 =

3( 1) + 4 2
s
s + O( 2 ),
+
au
6(a u)3

4 =

3
1
s + O(),
+

41

5 =

2
3
s + O(),
+

42

3(1)+4

6(a+u)
a

1

+ O( 2 )
= ( 1)T + s (1)T
a(a+u)

1
0

0
u



2 = T + s 0 + O( 2 )


1
0
3(1)4

6(au)

3

+ O( 2 )
= ( 1)T + s (1)T
a(au)

1
0


u
0


2
u
4 = (1)T
+
s

1 + O( )
1 u

0
1


u
0


4
u
+ s 1 + O( 2 )
= (1)T

u
2
0
1

where = /Pr and 1,2 satisfy

(24)

(25)

(26)

(27)

(28)

3 (u2 a2 /) + (4/3)u2
3 2
(u a2 ), 1 + 2 =
.
4
4
u
To date the only the zeroth-order form of the boundary conditions have been constructed, following Halpern. For
the case of non-reflecting subsonic outflow one has
!
T
u 1 a u
(29)
=
u

x
Z
a
T
!
T
1 aU
T

=
u
(30)
x
Z
a
T
1 2 =

where Z = u a(U/T 1) and U = ( 1)T u/(1 u). Observe that when = 0 one recovers the characteristic
condition that u/au T/T / = 0.

IV. Specific boundary conditions


To focus the general formulation presented in the previous sections on specific boundary types (walls, etc.) a
number of boundary conditions are derived based on Eq. (17) in a manner analogous to that proposed by Thompson.3, 4
Not all of the possible boundary conditions will be presented but a detailed derivation of a useful subset of the possible
conditions will be given. Additional boundary conditions can be easily derived in a similar manner. It is further stated
that when the grid is stationary the present boundary conditions resort to those found in Kim & Lee.12, 13
A.

Inviscid deformable wall

Consider the case where a non-rigidly moving wall bounds one portion of the fluid domain. A schematic of a -constant
wall is given in Figure 1, with unit normal n and wall velocity vector x wall . On this surface we require that
u n = x wall n

(31)

= (0, 0, 0, |c|, |c|). This fact results from the


be satisfied; we denote x wall n by vw . We observe that the diagonal of
conformally-deforming mesh which has metrics that satisfy t = vwall x = vw for all time such that U + t = 0.
Hence only the L4 element enters the domain and it must be chosen such that
U
= v w .

6 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

(32)

This requirement suggests that the replacement


L4 L5 2

(33)

be used.

Figure 1: Wall lying on a constant -plane.

B.

Viscous, isothermal deformable wall

As for the inviscid deformable wall we assume the geometry of Figure 1. For a viscous, isothermal deformable wall
we require that

u wall = x w
(34)

T wall = T w
(35)

where the terms on the right-hand-sides of the above equations represent the wall velocity and temperature, respectively. The wall density is updated according to the continutity equation
 

+
(36)
U j = 0.
J
j

The momentum and total energy density are updated using Eq. (36) according to

x w
(u) = x w
+

T w
1
x w

(E) = Cv T w + x w x w
+ Cv
+ x w

(37)
(38)

a calorically perfect gas with a specific heat at constant volume of Cv has been assumed.
C.

Inviscid subsonic inflow

In this application we take an inflow boundary to exist along a -plane and that the boundary does not move. This
latter assumption is not restrictive but is the form most commonly applied. Under these conditions we observe that
are positive, assuming U is strictly positive, implying we are to enforce four conditions to
the first four elements of
V,
and W,
along with the
estimate the amplitudes L1 , . . . , L4 . Choosing to specify the three velocity components U,
temperature T , following PL, suggests the following substitutions
L4 L5 2
L2
L3

(39)
(40)

( 1)
T
(L4 + L5 ) +
L1 +
2c
T
where L5 is computed using information from the domain interior.
7 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

(41)
(42)

V. Applications
In this section a series of example applications are presented to demonstrate the boundary conditions in various
contexts. In cases where possible the numerical data is compared to analytical solutions. These problems have been
chosen because the body motion is the sole source of unsteadiness and thus test specifically the ability of the boundary
conditions to handle deforming boundaries.
The first two examples consider moving cylinders placed in a quiescent environment whose motion is known to
generate acoustic waves.22 The third example presents a fluid-structure coupled problem of an initially displaced
circular elastic ring, emersed in a quiescent fluid, into which the ring radiates energy as sound.23 The fourth and
final example considers the one-dimensional nonlinear problem of an impulsively displaced piston in a constant area
channel. In the first three problems an analytical solution is available in the linear limit while a nonlinear analytical
solution, based on the method of characteristics, is available for the fourth problem until shock formation.
In all simulations the same numerical code is used which implements the boundary conditions formulated in
Sections II and III. The code was developed by the author and is based on the work of Visbal & Gaitonde.24 The
conservation equations of continuity, momentum, and energy are written in conservative form in the transformed space
and transformed time .17 The spatial derivatives use the sixth order standard compact finite difference stencil11 and
the temporal integration uses the standard fourth order Runge-Kutta scheme. For the boundary conditions the matrix
P and its inverse P1 of Eq. (15) are computed at each point along the boundary and the boundary value updates are
implemented as Eq. (17). No filtering of the variables is performed throughout the solution history.
A.

Translating cylinder

Consider a circular cylinder of fixed radius a and center (x0 (t), 0) whose time-dependent motion is constrained along
the x axis. The cylinder is immersed in a quiescent fluid of uniform density , uniform pressure p , and zero velocity.
If the translation in the x-direction is of the form
x0 (t) = w(t) sin t

(43)

where w(t) is a suitably smooth ramp function, such that w(0) = 0 and w(t ) = 1, then when a and c /a
the fluid response may be regarded as linear. If, in addition, we assume the fluid to be inviscid, then the linear wave
equation for the pressure
2 p
2 p
=0
(44)
a2
2
xi xi
t
holds with the boundary condition

p
v
x
= w r = r ,
(45)
r r=a
t
t
with r the unit outward radial normal, applied at r = a after linearization.
Together with the Sommerfeld radiation condition the time-harmonic solution of the above problem is easily found
using separation of variables and is given by
p(x,
) =

An () cos(n)Hn(1) (kr)

(46)

n=0

where k = /c is the acoustic wavenumber and Hn(1) is the Hankel function, of order n, of the first kind.25 For motion
only along the x axis only A1 is nonzero and is given by
A1 () =

2i c V c ()
(1)
H0 (ka) H2(1) (ka)

where
V c () =

x0 (t)eit dt

(47)

(48)

is the Fourier transform of the center velocity. For the present case of nonharmonic motion the pressure is given by
the inverse Fourier transform of Eq. (46) as
1
p(x, t) =
2

p(x,
)eit d.

8 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

(49)

A single block structured grid was used to describe the translating cylinder, as depicted in Fig. 2. The instantaneous
cylinder surface is located on the inner boundary Ri while the fixed, outer surface of radius 5a is at R0 . The mesh is
nonorthogonal throughout the domain: the origin of the radial lines of the mesh are offset vertically by an amount
y0 = a/2 from the instantaneous center. For a specific cylinder motion we take
w(t) =

1
[1 + tanh(5(t 2))]
2

=1

and

(50)

in Eq. (43) consider and two different amplitudes of = 0.05 and = 0.0005.
The resulting response taken at (x, y) = (x0 (t) + a, 0), i.e., at the point on the cylinder surface that lies on the x
axis, is shown in Fig. 3. For a purely linear response all three curves shown in the figure should collapse on to a single
curve. This is approximately the case but one can discern that the = 0.0005 response is closer to the analytical
response than is the high amplitude motion, indicating some presence of nonlinearity.

y/a

y0

Ri

-2

Ro

-4

-6

-4

-2

x/a

Figure 2: Non-orthogonal mesh for translating and breathing cylinders.

[(p p )/p]/

0.5

0.5

1.5
0

ta/a

Figure 3: Pressure at (x, y) = (x0 (t) + a, 0) for the translating cylinder. Legend: , analytical solution; ,
= 0.05; , = 0.0005.

9 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

B.

Breathing cylinder

In this section we allow the radius of the cylinder to be a function of time and keep the center fixed at (0, 0). The radius
at time t is given by
r(t) = a(1 + w(t) cos t)
(51)
with w(t) and as in Eq. (50) and again consider two amplitudes = 0.05 and = 0.0005. In a manner similar to that
given in IV. A. the analytical pressure response is determined by the inverse Fourier transform of Eq. (49) but with
p(x,
) =
for the pressure Fourier amplitude and
V r () =

i c V r
H1(1) (ka)
Z

H0(1) (kr)

(52)

r(t)eit dt

(53)

for the radial velocity Fourier transform.


Using the same single block structured grid shown in Fig. 2 the results for the pressure on the cylinder surface at
(x, y) = (a(t), 0) are shown in Fig. 4. Again we note that in the purely linear case all three curves should collapse upon
a single curve and that the simulation data approximately demonstrate this behavior, with the = 0.0005 being the
closer of the two simulations to the anaytical solution.
2

[(p p )/p]/

1.5
1
0.5
0

0.5
1
0

tc /a

Figure 4: Pressure at (x, y) = (a(t), 0) for the breathing cylinder. Legend: , analytical solution; , = 0.05;
, = 0.0005.

C.

Fluid-structure interaction

A slightly more interesting validation case concerns the coupled motion of a linearly elastic thin cylindrical shell in a
quiescent medium. Consider the situation when the cylinder, which has a mean radius a at equilibrium, is symmetrically displaced in the outward radial direction by an amount w0 as shown in Fig. 5. Under the restriction of h/a 1
and axisymmetric displacement the motion of the solid is described by linear, second order ordinary differential equation whose solution depends on pressure load distribution. The pressure likewise depends on the acceleration of the
cylinders surface as seen in the previous two examples.
As before the requirement that w0 a lends the problem to linearization and, hence, to analytical treatment.
Because the problem of interest is an initial boundary value problem the Laplace transform is used in lieu of the
Fourier transform. Details of the solution are provided in the Appendix; the relvant detail is that the early response of
the cylinder displacement, w(t), from the equilibrium condition is given by the simple expression
w(t) = exp{t/2} cos t

10 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

(54)

where = [(1 )2 c ]/[(1 2)h s ], = Cph /a, is Poissons ratio, h is the shell thickness (see Fig. 5), s is the
p
density of the solid, and Cph = E/[(1 2 ) s ] is the wavespeed of the solid. The rate of energy transmission from
the solid to the fluid is proportional to = /(2) and is a fundamental quantity in the structural-acoustic response of
the system.
The specific case under consideration is a thin circular shell of thickness h/a = 0.01, a Poissons ratio of = 0
for plane strain, a density ratio of s / = 2700, and a phase speed ratio of Cph /c = 17.9. These values roughly
correspond to an aluminum cylinder of 1 m radius. The structure will vibrate with a frequency of = 970 Hz and has
a damping ratio of = 0.00134. Figure 6 shows the agreement between the numerical and the analytical solution for
the displacement history of the cylinder. (The analytical solution involves a numerically inverted Laplace transform
of Eq. (60) derived in the Appendix.) It is apparent that the damping rate of the two solutions is quite similar. The
numerical estimate of the damping rate is subject to uncertainities as there are a variety of methods to determine it.
Using the peak-to-peak decrement in the w(t) history shown in Fig. 6 gives num = 0.013 0.005, which is consistent
with the expected value.

Displaced shell at t = 0

, c

1
0.8
0.6
0.4
0.2
w
0
w0
0.2
0.4
0.6
0.8

0.02

w0

0.04

0.06

Figure 5: Coupled fluid-structure motion of a


radially-displaced thin-shelled cylinder.

Figure 6: Vibration response of a thin-shelled cylinder immersed in a fluid. Legend: , analytical


solution; , numerical solution.

VI. Piston motion within a channel


The last example problem to be presented is that of an accelerating piston enclosed in a one-dimensional channel.
As shown in Fig. 7 a tighly fitted piston in a constant area channel (of diameter D) is moves to the right with position
history X(t). Before the time any shock forms one can use the method of characteristics to determine the flow properties
as a function of x and t. For a fixed position x0 the density may be expressed as a simple function of the piston motion.
When the channel is filled initially with a quiescent fluid of density 0 and pressure p0 is straightfoward to show
that the density at a fixed point x0 is given by the piecewise expression

t x0 /c0
0
(t; x0 ) =
(55)
h 2 i1/(1)

c
t x0 /c0
s0

where s0 = p0
0 and c(t) = [(1)/2] X()+c0 where (t) is given implicitly by x0 = X()+{[(+1)/2] X()+c0 }(t).
2
The results from the specific case of X(t) = [A/2]t , A = 1/10 and x0 = 6 are given in Fig. 8 where the agreement
between the two solutions is good.

11 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

2
1.8

/0

1.6
1.4
1.2

Piston
1

X(t)
D

0.8
0

tc0 /D

Figure 8: Density response at a point due to piston


motion. Legend: , analytical; , numerical.

Figure 7: Motion of piston through a onedimensional channel.

VII. Conclusions
An extension to grids of arbitrary deformation of the characteristics-based boundary conditions was proposed. By
formulating the problem in curviliner coordinates with a time-dependent metric inclusion of the grid motion into the
boundary conditions is straight forward and closely follows the structural form developed by Thompson3, 4 and Poinsot
& Lele.5 When the grid is rigid the present boundary conditions naturally resort to those given by Kim & Lee.12, 13 For
viscous problems the well posed method of Halpern16 was presented and carried out for the case of a subsonic viscous
inflow for a zeroth order condition; further work is needed to derive the 1st order boundary condition.
Verification of the boundary conditions by five example problems with analytical solutions shows their utility
in various cases, including linearized acoustics, fully coupled fluid-structure interactions, and nonlinear waves. The
simulations highlight the accuracy of the boundary formulation and did not exhibit numerical instability. However,
since we do not have proofs of stability, via an energy norm2 or otherwise, we cannot claim these boundary condtions
are strictly stable in all cases.

A. Analytical solution of the vibration of thin-shell cylinder in a fluid


A derivation of the solution for the coupled response of a thin-shelled cylinder immersed in a quiescent fluid is
given here. Under the assumptions listed in IV.C the governing equation for the solid is23
2
Cph

d2 w w
(1 2 )(p p )
+
=
Eh
dt2
a2

(56)

where E is the materials Youngs modulus and we consider only the influence of the fluid outside of the cylinder. The
fluid is governed by the usual inviscid wave equation given in Eq. (44). These equations are supplemented with the
initial and boundary conditions
w(0) = w0 , p p

d w p
=0
2 +
r r=a
dt
!
p
1 p
lim r
+
= 0.
r
r a t

at t = 0

(57)

The last condition given is the Sommerfeld radiation condition for a two-dimensional acoustic medium.
On using the Laplace transform
Z

f (t) exp{st} dt
f (s) =
0

12 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

(58)
(59)

the Sommerfeld condition immediately implies that p(x,


s) = A(s)H0(2) (isr/a ) where A(s) is an unknown function
to be determined from the remaining equations and boundary conditions. Upon combining the Laplace-transformed
equations Eq. (56) and Eq. (44) and the boundary conditions, we find that the Laplace-transformed radial displacement
of the solid is given by
w(s)

=
w0

s
2
Cph
s2
2
Cph

1
a2

where = isa/c .
The inverse Laplace transform
f (t) =

1
2i

H (2) ()

+ i (1Eh) c H0(2) ()
1

2
(12 ) c Cph
H0(2) ()
is
Eh
H1(2) ()

(60)

f(s) exp{st} ds

involves the Bromwich contour which must lie to the right of any poles of f(s). It is straightforward to verify that
Eq. (60) does not have any poles in the right half of the complex s-plane thus we can deform to the imaginary s axis.
At present we do not have an analytical form of the time domain expression of the radial displacement w(t), aside from
the inverse Laplace transform, which much be performed numerically. However, the inverse Laplace transform of the
limit of w(s)

as s gives an estimate to the early time response, which is instructive. In taking the s limit
we find that Eq. (60) simplifies to (using the expressions for and beta below Eq. (54))
s+
w(s)

= 2
w0
s + s + 2
which has the simple inverse Laplace transform of
s
s
!2
!2

w(s)

/(2)

(/2)t
+ q
.
t 1
= e 2 cos t 1
e
sin
 2

w0
2
2

1 2

(61)

(62)

For the 1 m Aluminum shell paramaters given in IV. C we have that / 1 such that the approximate early time
response is given by Eq. (54). Note that the analytical curve in Fig. 6 uses the numerically calculated inverse Laplace
transform of Eq. (60).

References
1 Oliger, J. and Sundstr
om, A., Theoretical and Practical Aspects of Some Initial Boundary Value Problems in Fluid Dynamics, SIAM J.
Appl. Math., Vol. 35, 1978, pp. 419446.
2 Nordst
om, J. and Svard, M., Well-posed boundary conditions for the Navier-Stokes equations, SIAM J. Numer. Anal., Vol. 43, No. 3, 2005,
pp. 12311255.
3 Thompson, K. W., Time Dependent Boundary Conditions for Hyperbolic Systems, J. Comp. Phys., Vol. 68, 1987, pp. 124.
4 Thompson, K. W., Time Dependent Boundary Conditions for Hyperbolic Systems, II, J. Comp. Phys., Vol. 89, 1990, pp. 439461.
5 Poinsot, T. J. and Lele, S. K., Boundary Conditions for Direct Simulations of Compressible Viscous Flows, J. Comp. Phys., Vol. 101, 1992,
pp. 104129.
6 Kreiss, H., Initial boundary value problems for hyperbolic systems, Commun. Pure Appl. Math, Vol. 23, No. 3, 1970, pp. 277298.
7 Hesthaven, J. S. and Gottlieb, D., A stable penalty method for the compressible Navier-Stokes equations. I. Open boundary conditions,
SIAM J. Sci. Comput., Vol. 17, No. 3, 1996, pp. 579612.
8
Svard, M., Carpenter, M. H., and Nordstrom, J., A stable high-order finite difference scheme for the compressible Navier-Stokes equations,
far-field boundary conditions, J. Comput. Phys., Vol. 225, 2007, pp. 10201038.
9
Svard, M. and Nordstrom, J., A stable high-order finite difference scheme for the compresible Navier-Stokes equations: No-slip wall
boundary conditions, J. Comput. Phys., Vol. 227, 2008, pp. 48054824.
10 Strand, B., Summation by parts for finite difference approximations for d/dx, J. Comput. Phys., Vol. 110, 1994, pp. 4767.
11 Lele, S. K., Compact Finite Difference Schemes with Spectral-like Resolution, J. Comp. Phys., Vol. 103, 1992, pp. 1642.
12 Kim, J. W. and Lee, D. J., Generalized characteristic boundary conditions for computational aeroacoustics, AIAA J., Vol. 38, No. 11, 2000,
pp. 20402049.
13 Kim, J. W. and Lee, D. J., Generalized characteristic boundary conditions for computational aeroacoustics, Part 2, AIAA J., Vol. 42, No. 1,
2004, pp. 4755.
14 Engquist, E. and Majda, A., Absorbing boundary conditions for the numerical simulation of waves, Math. Comput., Vol. 31, 1977, pp. 629
651.
15 Rudy, D. and Strikwerda, J., A nonreflecting outflow boundary condition for subsonic Navier-Stokes Calculations, J. Comp. Phys., Vol. 36,
1980, pp. 6670.

13 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

16 Halpern, L., Artificial boundary conditions for incompletely parabolic perturbations of hyperbolic systems, SIAM J. Math. Analy., Vol. 22,
No. 5, 1991, pp. 12561283.
17 Vinokur, M., Conservation equations of gasdynamics in curvilinear coordinate systems, J. Comp. Phys., Vol. 14, 1974, pp. 105125.
18 Hirsch, C., Numerical Compuation of Internal and External Flows, Volume 2, John Wiley & Sons, Inc., New York, NY, 1990.
19 Loh
eac, J.-P., An Artificial Boundary Condition for an Advection-Diffusion Equation, Math. Methods Appl. Sci., Vol. 14, 1991, pp. 155
175.
20 Sutherland, J. C. and Kennedy, C. A., Improved boundary conditions for viscous reacting compressible flows, J. Comput. Phys., Vol. 191,
2003, pp. 502524.
21
Engquist, B. and Majda, A., Radiation boundary conditions for acoustic and elastic wave calculation, Comm. Pure & Appl. Math, Vol. 32,
1979, pp. 313357.
22
Morse, P. M. and Ingard, K. U., Theoretical Acoustics, Princeton University Press, Princeton, NJ, U.S.A., 1968.
23 Junger, M. C. and Feit, D., Sound, Structures, and Their Interaction, Acoustical Society of America, Melville, NY, U.S.A., 1993.
24 Visbal, M. R. and Gaitonde, D. V., On the Use of Higher-Order Finite-Difference Schemes on Curvilinear and Deforming Meshes, J.
Comp. Phys., Vol. 181, 2002, pp. 155185.
25 Abramowitz, M. and Stegun, I., editors, Handbook of Mathematical Functions, Dover Publications, Inc., New York, U. S. A., 1968.

14 of 14
American Institute of Aeronautics and Astronautics Paper 2009-0010

Das könnte Ihnen auch gefallen