Download as pdf or txt
Download as pdf or txt
You are on page 1of 15

Journal of Sound and Vibration (1996) 193(5), 1099–1113

ATTENUATION OF HIGHER ORDER


CIRCUMFERENTIAL THERMOACOUSTIC
WAVES IN VISCOUS FLUID LINES
P. N. L†
Aeroacoustic Corporation, 169 Highland Parkway, Roselle, NJ 07203, U.S.A.

H. A. S
Laboratory for Noise and Vibration Control Research, Department of Mechanical
Engineering, Aeronautical Engineering and Mechanics, Rensselaer Polytechnic Institute,
Troy, NY 12180-3590, U.S.A.

(Received 20 March 1995, and in final form 31 October 1995)

The acoustic waves propagation in viscous water, glycerin and air contained in a rigid
wall, thermally insulated, infinite long, circular tube are studied using the exact
three-dimensional thermal–fluid coupled equations for the vibrations in the n = 0, 1
circumferential modes. The first three axially symmetric modes at n = 0 and the first three
non-axially symmetric modes at n = 1 are presented. The corresponding two-dimensional
mode shapes are plotted so that the wave vibrations can be identified. It is found that the
dispersion spectra, mode shapes and phase velocity dispersion plots of three fluid mediums
are very close. But the attenuation rates of glycerin and air are about 37 and 9 times higher
than water, respectively.
7 1996 Academic Press Limited

1. INTRODUCTION
The attenuation of sound propagation in fluids comes from two paths. One is from the
‘‘volume absorption’’ due to vibration energy converting to heat in the fluid itself through
viscous and thermal conducting processes. The other is from the absorptive walls through
surface impedance.
The applications of the first type of absorption are air absorption in rigid walls for long
ventilation ducts or viscous dissipation in liquid-filled pipes. The relevant equations of
motion are the viscous–compressible Navier–Stokes equation for the fluid system. This
case was first studied by Kirchhoff [1], then was studied by Gerlach and Parker [2] and
Scarton and Rouleau [3]. Later, the thermal effect was studied by Tijdeman [4] and Huerre
and Karamcheti [5].
The applications of the second type of absorption are sound attenuation in lined ducts
for air ventilation systems or silencers for noisy fans. Since wall absorption is much more
significant than volume absorption in this type of application, the viscous and thermal
effects are normally neglected for simplicity. This case has been studied by Zwikker and
Kosten [6], Doak and Vaidya [7], Munjal [8] and Liang [9].

†Now with Harley-Davidson Motor Co., NVH Engineering, 3700 West Juneau Ave., Milwaukee, WI, 53201,
U.S.A.
1099
0022–460X/96/251099 + 15 $18.00/0 7 1996 Academic Press Limited
1100 . .   . . 
Since fluid system equations derived from the volume absorption are more accurate (but
complex), this approach may also be useful for wall absorption problems, if moderately
modified. For instance, the model of sound propagating in silencers can be treated
separately as sound propagating in an air slot and sound propagating in the fluid acoustic
fill with coupling occurring by using boundary conditions at the interface.
The theoretical frame work for the thermal effect used in this paper was based on a
suggestion by Doak [10] and the work of Huerre and Karamcheti [5]. Monkewitz [11] has
provided the essential numerical techniques to solve the mathematical algebra in the terms
of vector potential function in the velocity field. Chang [12] helped to start the first phase
of this research.
Doak [10] suggested that the fluctuating velocity field should be considered as the
superposition of acoustic (related to pressure), thermal (related to entropy) and viscous
(related to vorticity) parts. Huerre and Karamcheti [5] analyzed the friction and heat
conduction effects on sound propagation in rectangular ducts for air. This research will
extend the above work to an insulated, rigid, impermeable circular duct applied to specific
fluids (two liquids, water and glycerin, and one gas, air).

2. THEORY
For small amplitude laminar motions of a simple thermodynamic Newtonian fluid
without net flow, all non-linear terms in the perturbation variables are neglected. The
following linear thermoacoustic equations result: continuity equation,
1t r + r0 (9 · v) = 0, (1)
Navier–Stokes equation,
1
1t v = − 9P − n0 9 × (9 × v) + (n'0 + 43 n0 )9(9 · v), (2)
r0
thermodynamic energy equation,
r0 T0 1t s = kf 9 2T. (3)
Two modified state equations are cited from Huerre and Karamcheti [5]:

6 0 1 7
1/2
1 (g0 − 1)T0
r= P − r0 C 0 s (4)
C02 cp

and

0 1
g0 T0
1/2
C0 (g0 − 1)T0
T= r+ s. (5)
r0 cp cp

The subscript zero represents the initial state of the medium. The five fluctuating quantities
r, v, s, T and P are similar in form. For instance, the density fluctuation is r = rtotal − r0 ,
kg/m3. The units for velocity fluctuation v, entropy fluctucation s, temperature fluctuation
T and pressure fluctuation P are m/s, J/°K, °K and N/m2, respectively.
The remaining constants are similarly defined: n0 is the first coefficient of kinematic
viscosity, n0 = m0 /r0 in m2/s; n'0 is the second coefficient of kinematic viscosity in m2/s. In
the low frequency range of interest, n'0 is approximately equal to zero [3], which is known
as Stoke’s assumption; kf is the thermal conductivity of fluid in W/(m °K); C0 is the intrinsic
sound speed in the fluid, m/s; g0 is the ratio of specific heats, g0 = cp /cv , where cp and cv
are the specific heat at constant pressure and at constant volume, respectively, in J/(kg °K).
  1101
Using the Helmholtz decomposition theorem, the velocity fluctuation v can be expressed
by the sum of three distinct terms:
v = 9(fa + fth ) + 9 × C, (6)
where the scalar potential f has been divided into the sum of the acoustic potential fa
and the thermal potential fth as per Doak’s suggestion [10] and C is the vector potential
caused by viscous effects [3].
The continuity equation (1) can be expressed as
1t r = −r0 9 2(fa + fth ). (7)
Two very useful thermodynamic quantities 1t P and 1t s can be obtained by taking the
derivative of the modified state equation (4) with respect to time t, and then by using the
continuity equation (7),

0 1
r (g0 − 1)T0
1/2
1
−r0 9 2(fa + fth ) = 1 P− 0 1t s. (8)
C02 t C0 cp

We can arrange the acoustic potential fa and the thermal potential fth so that one
corresponds to pressure fluctuation 1t P, and the other corresponds to entropy fluctuation
1t s as follows:

0 1
1/2
cp
1t P = −r0 C02 9 2fa , 1t s = C 0 9 2fth . (9, 10)
(g0 − 1)T0

Using the Helmholtz decomposition theorem, equation (6), and imposing the Lorentz
gauge condition 9 · C = 0 to the Navier–Stokes equation (2), gives
1
9[1t (fa + fth )] + 9 × 1t C = − 9P + n0 9 × 9 2C + (n'0 + 43 n0 )9[9 2(fa + fth )]. (11)
r0
For the decomposition of the scalar potentials f and the vector potential C in the above
equation (11), one takes the divergence and curl, respectively, to give
P
1t (fa + fth ) = − + (n'0 + 43 n0 )9 2(fa + fth ), (12)
r0
1t C = n0 9 2C. (13)
Taking the derivative of equation (12) with respect to time t and then replacing the quantity
1t P by using equation (9),
C02 9 2fa − 1t2 fa + (n'0 + 43 n0 )9 2(1t fa ) = 1t2 fth − (n'0 + 43 n0 )9 2(1t fth ). (14)
Equation (14) can be expressed succinctly as
1
q2fa = 1 (1 f − (n'0 + 43 n0 )9 2fth ), (15)
C02 t t th
where the wave operator is defined as

0 1
1 1
q2 = 1 + 1 9 2 − 2 1t2 , (16)
v0 t C0
1102 . .   . . 
and v0 (rad/s) is defined as
v0 = C02 /(n'0 + 43 n0 ). (17)
Substituting the quantity 1t s from equation (10) into the energy equation (3) leads to

0 1
1/2
cp
r0 T0 C0 9 2fth = kf 9 2T. (18)
(g0 − 1)T0

The above equation implies

0 1
n0 (g0 − 1)cp
1/2

fth = T, (19)
Pr C0 T0

where the Prandtl number Pr is defined as Pr = r0 n0 cp /kf .


Taking the derivative of equation (19) with respect to time t gives

0 1
1/2
Pr C0 T0
1t T = 1t fth . (20)
n0 (g0 − 1)cp

Also, taking the derivative of the modified state equation (5) with respect to time t, and
utilizing the continuity equation (7) and the quantity 1t s in equation (10) gives

0 1
1/2
T0
1 t T = C0 9 2[fth − (g0 − 1)fa ]. (21)
(g0 − 1)cp

Comparing the above two equations (20, 21), one can immediately obtain the relationship
n0 2 n
−(g0 − 1) 9 fa = 1t fth − 0 9 2fth . (22)
Pr Pr
Equations (13), (15) and (22) constitute the fluid system equations. Furthermore,
cross-differentiation of equations (15) and (22) can decouple the thermal and acoustic
terms and leads to the additional fourth order partial differential equation
Pr
9 2(15) = q2(22). (23)
(1 − g0 )n0
After careful manipulation of the above equation, one can obtain the decoupled fluid
system equations as
n0 2 2 n0 2 2
1t q2fa = 9 qt fa , 1t q2fth = 9 qt fth , 1t C = n0 9 2C, (24–26)
Pr Pr
where the thermal wave operator q2t is defined as

0 1
g0 g
q2t = 1 + 1 9 2 − 02 1t2 , (27)
v0 t C0

and where v0 is defined by equation (17).


From the inspection of equations (24) and (25), one can find the relationship
fth = dfa , (28)
  1103
where the fluid thermal coupling factor, d, is a function of the frequency v and will be
defined later by equation (37).
Assuming all quantities to be a simple harmonic function of u, z and t for the scalar
potential f and vector potential C, the potential functions can be expressed as
fa (r, u, z, t) = f
a (r) einu + kz + ivt, C(r, u, z, t) = C
(r) einu + kz + ivt, (29, 30)
where i = (−1)1/2, v is the radian frequency, and k = kr + iki is the complex axial
wavenumber. The real part is the attenuation constant and the imaginary part is the
propagation constant. Expansion and rearrangement of the system equation (24) leads to

6 00 1 1> 7
v2 Pr v 2 P v2 P v
94 + 2 g0 + − iv r 9 2 − iv 2 r (1 + i g ) f
a (r) einu + kz = 0. (31)
C0 n0 v0 n0 C 0 n0 v0 0

Equation (31) can be rewritten as


(9 2 + h12)(9 2 + h22)f
a (r) einu + kz = 0, (32)
where

60 1
1 v2 P v2 P
2
h1,2 = g + r − iv r
2 C02 0 n0 v0 n0

00 1 1 7>0 1
v2 Pr v 2 v 2 Pr v
2 1/2
P
2 2 g0 − + iv r − 4iv (g − 1) 1+i g . (33)
C0 n0 v0 n0 C02 n0 0 v0 0

Since two harmonic operators in equation (32) are obviously commutative, the resulting
soltuion (34) can easily be shown to be satisfied by an combination of the two Helmholtz
equations (32):
f
a,n (r) = C1 Jn (g1 r) + C2 Jn (g2 r), (34)
where
gj2 = hj2 + k 2, j = 1, 2. (35)
The relationship between the thermal potential f
th and the acoustic potential f
a is
obtained by substituting 9 2fa = −hj2 fa and 9 2fth = −hj2 fth into equation (22):
(fth )j = dj (fa )j or (f
th )j = dj (f
a )j , j = 1, 2, (36a, b)
where dj is defined as the fluid thermal coupling factor
dj = (g0 − 1)/[1 + iv(Pr /n0 hj2 )], j = 1, 2. (37)
The scalar potential f
is the superposition of two scalar potentials f
a and f
th :
f
= f
a + f
th = (1 + dj )f
a j = 1, 2, (38a)
or
f
n (r) = (1 + d1 )C1 Jn (g1 r) + (1 + d2 )C2 Jn (g2 r). (38b)
There is a special case in the fluid system: if the ratio of specific heat g0 = 1 (for example
water, g0 = 1·004 in Table 1), then the fluid is said to be barotropic. The thermal coupling
factor d is equal to zero. The thermal potential fth , therefore, disappears. The system
1104 . .   . . 
equation (15) reduces q fa = 0. This case was extensively studied by Scarton and Rouleau
2

[3]. Meanwhile, two parameters h12 and h22 can be reduced to

>0 1
v2 v Pr
h12 = 1+i g , h22 = −iv . (39a, b)
C02 v0 0 n0

The solution of system equation (26) for the vector potential function can be shown,
after some manipulations, to be

C
n (r) = C
r,n (r)er + C
z,n (r)ez + C
u,n (r)eu , (40)

where

0 1
n n
C
r,n (r) = Q · i J (lr) + S · i Jn + 1 (lr) − Jn (lr) , (41)
lr n lr

l
C
z,n (r) = −S · i J (lr), (42)
k n

0 1
n n
C
u,n (r) = Q Jn + 1 (lr) − J (lr) + S · Jn (lr), (43)
lr n lr

and where,

l 2 = k 2 − i(v/n0 ). (44)

Now the components of the velocity v in equation (6) can be expressed by a series of
Bessel functions through f and C:

1
vx r,n (r) = 1r f
n + 1u C
z,n − 1z C
u,n
r

1
= {C1 (1 + d1 )(nJn (g1 r) − g1 rJn + 1 (g1 r)) + C2 (1 + d2 )(Jn (g2 r) − g2 rJn + 1 (g2 r))
r

0 1
nk v
+Q J (lr) − krJn + 1 (lr) −S.i.n J (lr)}, (45)
l n lkn0 n

T 1
Thermal properties of water, glycerin and air at 293 °K (20°C or 68°F). The radius ‘‘a’’ of
fluid line is 0·0414m (1·63in) and the second coefficient of viscosity m'0 is set to 0
Water Glycerin Air
Sound speed C0 , m/s 1481 1895 343
Shear viscosity m0 , kg/(m.s) 0·001 1·49 0·0000181
Density r0 , kg/m3 998 1228 1·21
Ratio of specific heats g0 1·004 1·117 1·402
Const. press. spec. heat cp , J/(kg°K) 4181 2340 1005·6
Thermal conduct. of fluid kf , W/(m°K) 0·556 0·2789 0·02568
Prandtl number Pr 7·520 12500 0·7088
  1105
1 1
vx z,n (r) = 1z f
n + 1r C
u,n + C
u,n − 1u C
r,n
r r
= C1 (1 + d1 )kJn (g1 r) + C2 (1 + d2 )kJn (g2 r) + QlJn (lr), (46)
1
vx u,n (r) = 1u f
n + 1z C
r,n − 1r C
z,n
r

6
1 k
= C1 .i.n(1 + d1 )Jn (g1 r) + C2 .i.n(1 + d2 )Jn (g2 r) + Q.i.n Jn (lr)
r l

0 17
v 1
+S nJ (lr) − rJn + 1 (lr) . (47)
kn0 l n

The complete form of velocity v is expressed as


a
v = s (v̂r,n (r)er + v̂z,n (r)ez + v̂u,n (r)eu )einu + kz + ivt. (48)
n=0

The temperature T
can be expressed in terms of f
th through equation (19) as

0 1
1/2
P r C0 T0
T
= f
th , (49)
n0 (g0 − 1)cp

where
f
th = d1 (f
a )1 + d2 (f
a )2

0 1
1/2
Pr C0 T0
= {C1 d1 Jn (g1 r) + C2 d2 Jn (g2 r)}. (50)
n0 (g0 − 1)cp

The gradient of T
in the radial direction can be derived as

0 1
1/2
Pr C0 T0 1
1r T
= {C1 d1 (nJn (g1 r) − g1 rJn + 1 (g1 r))
n0 (g0 − 1)cp r

+ C2 d2 (nJn (g2 r) − g2 rJn + 1 (g2 r))}. (51)


The fluid radial normal stress trr , axial shear stress trz and circumferential shear stress
tru are expressed as follows. The detailed derivatives are listed in reference [13].

0 1 0 1
i 1 1
t̂rr,n (r) = − r0 C02 9 2f
a + P0 + 2n0 r0 1r2 f
+ 1r 1u C
z − 2 1u C
z − 1r 1z C
u
v r r

+ (n'0 − 23 n0 )r0 9 2f

6 0 0 1 1
n 0 r0 n' 2
= C1 (1 + d1 ) 2(n 2 − n − g12 r 2 )− 0 − h12 r 2 Jn (g1 r)
r2 n0 3

0 1
C02
+ C1 i.h12 r 2 · J (g r) + 2g1 rJn + 1 (g1 r)
n0 v n 1
1106 . .   . . 

0 0 1 1
n'0 2 2 2
+ C2 (1 + d2 ) 2(n 2 − n − g22 r 2 )− − h r Jn (g2 r)
n0 3 2

0 1
C02
+ C2 i · h22 r 2 J (g r) + 2g2 rJn + 1 (g2 r)
n0 v n 2

0 1
k 2
+Q 2 (n − n − l 2r 2 )Jn (lr) + 2k · rJn + 1 (lr)
l

0 17
v v
+ S −2i(n 2 − n) J (lr) + 2i.n · r J (lr) , (52)
n0 lk n n0 k n + 1

0 1
iv 1
t̂rz,n = 2n0 r0 1r 1z f − n0 r0 21z2 − C
u + 2n0 r0 1z 1u C
z
n0 r

6
n0 r0
= C1 2k(1 + d1 )(nJn (g1 r) − g1 rJn + 1 (g1 r))
r

+ C2 2k(1 + d2 )(nJn (g2 r) − g2 rJn + 1 (g2 r))

0 1 7
1 v
+ Q(l 2 + k 2 ) nJ (lr) − rJn + 1 (lr) −S.i.n J (lr) , (53)
l n n0 l n

0 0 1 1
1 1 1 r iv
t̂ru,n = 2n0 r0 (1 − )1 f
− 1z C
r − 1u 1z C
u + 1u2 + 1r − C
z
r r r u r 2 n0

6
n0 r0
=2 C1 .i.n(1 + d1 )((n − 1)Jn (g1 r) − g1 rJn + 1 (g1 r))
r2

+ C2 .i.n(1 + d2 )((n − 1)Jn (g2 r) − g2 rJn + 1 (g2 r))

k
+ Q.i.n ((n − 1)Jn (lr) − lrJn + 1 (lr))
l

00 1 17
v 1 2 l 2r 2
+S n −n− Jn (lr) + rJn + 1 (lr) . (54)
n0 k l 2

3. BOUNDARY CONDITIONS AND CHARACTERISTIC EQUATION


It is assumed that the fluid column is contained in a rigid circular tube. The tube is well
insulated so that the heat transfer rate is zero at the wall (i.e. r = a in Figure 1):

vr = vz = vu = 0, 1r T = 0 (at r = a). (55)


  1107

Figure 1. The geometry and co-ordinate system of a rigid impermeable wall with thermally insulated, circular
tube of radius a = 0·0414m (1·63in).

Imposing four boundary conditions (55) into equations (45)–(47) and (51) results in the
matrix form
D · L = 0, (56)
where L = 6C1 , C2 , Q, S7 and D is a 4 × 4 matrix.
T

For non-trivial solutions of L, we should impose the determinant of matrix D equal to


zero and that leads to the characteristic equation,
det[Dj,k ] = 0, j(or k) = 1, 2, 3, 4. (57)
In order to simplify the problem, the dimensionless variables should be used in the elements
of the determinant which are defined as
z = ka, R = r/a, G1 = h1 a, G2 = h2 a, W1 = g1 a,

T 2
Comparison of thermal effect solutions of dimensionless complex axial wave numbers z at
the first two modes P10 and P11 in water, glycerin and air lines, where z = (zr ,zi ). The real
and imaginary parts are the attenuation and propagation constants, respectively. The
propagation constants, zi , of three mediums are very close. (a) Fundamental mode P10; (b)
first non-axially symmetric mode P11
(a)
Nondimensional
frequency Water Glycerin Air
0·01 (0·000009, 0·01001) (0·000294, 0·01028) (0·000074, 0·01007)
1 (0·000090, 1·00009) (0·002811, 1·00278) (0·000728, 1·00073)
4 (0·000181, 4·00018) (0·005805, 4·00556) (0·001473, 4·00145)
8 (0·000257, 8·00026) (0·008792, 8·00786) (0·002135, 8·00205)
12 (0·000315, 12·0003) (0·011703, 12·0096) (0·002697, 12·0025)

(b)
Nondimensional
frequency Water Glycerin Air
1·85 (0·000539, 0·18093) (0·015757, 0·19692) (0·004281, 0·18473)
2 (0·000187, 0·78124) (0·005886, 0·78680) (0·001513, 0·78255)
4 (0·000246, 3·55131) (0·007830, 3·55862) (0·001997, 3·55303)
8 (0·000360, 7·78560) (0·011987, 7·79628) (0·002964, 7·78813)
12 (0·000444, 11·8584) (0·015687, 11·8715) (0·003734, 11·8615)
1108 . .   . . 

Figure 2. Dimensional frequency V versus propagation constant zi and attenuation constant zr in water line.
All modes are located very near to either the pure imaginary or pure real surfaces. No complex modes are found
in the fluid system in the frequency range of interest: ——, axially symmetric modes n=0; - - - -, non-axially
symmetric mode n = 1; —w—, P10; - -w- -, P11; —r—, P20; - -r- -, P21; —q—, P30; - -q- -, P31; - -W- -,
A11; —R—, A20; - -R- -, A21; —Q—, A30; - -Q- -, A31.

W2 = g2 a, W3 = la,
V = (va)/C0 , D0 = n0 /(C0 a), D'0 = n'0 /(C0 a). (58)
The accuracy of the dimensionless eigenvalues z can be improved if one can improve the
matrix (i.e. reduce the condition number of determinant Dj,k ) by properly scaling the
system matrix. The fourth column in the matrix D should be scaled as
[D0 z/(iV)] · {Dj,4 } = {[D0 z/(iV)] · Dj,4 }, j = 1, 2, 3, 4. (59)
The resultant expressions of the dimensionless elements of Dj,k are as follows. For the three
velocity boundary conditions, v̂*r (R = 1) = v̂*
z (R = 1) = v̂*
u (R = 1) = 0:

D1,1 = −(1 + d1 ) F J
fnJn (W1 R)−W1 RJn + 1 (W1 R)j,
D1,2 = −(1 + d2 )(nJn (W2 R)−W2 RJn + 1 (W2 R)),
D1,3 = −F J
f(nz/W3 )Jn (W3 R) − zRJn + 1 (W3 R)j, D1,4 = (n/W3 )Jn (W3 R),

Figure 3. Dispersion of phase velocities in water line. The fundamental mode P10 is non-dispersive: ——,
axially symmetric modes, n = 0; ...., non-axially symmetric mode, n = 1; —w—, P10; - -w- -, P11; —r—, P20;
- -r- -, P21; —q—, P30; - -q- -, P31.
  1109

Figure 4. One-dimensional normalized mode shapes in radial direction r for water line at V = 4, n = 0, mode
P20, including real and imaginary parts. The radial and axial displacements at the fluid cylindrical surface r = a
are zero. No components of circumferential displacement and stress are presented in the axially symmetric mode,
n = 0: w, real part; r, imaginary part. (a) Radial displacement, ûr ; (b) axial displacement, ûz ; (c) radial normal
stress, t̂rr ; (d) axial shear stress, t̂rz ; (e) temperature, T
.

D2,1 = −(1 + d1 )zJn (W1 R), D2,2 = −(1 + d2 )zJn (W2 R),
D2,3 = −W3 Jn (W3 R), D2,4 = 0
D3,1 = −n(1 + d1 )Jn (W1 R), D3,2 = −n(1 + d2 )Jn (W2 R),
D3,3 = −(nz/W3 )Jn (W3 R), D3,4 = (n/W3 )Jn (W3 R)−RJn + 1 (W3 R),
For the thermal boundary condition, 1r T
*(R = 1) = 0:
D4,1 = −d1 (nJn (W1 R)−W1 RJn + 1 (W1 R)), D4,2 = −d2 (nJn (W2 R)−W2 RJn + 1 (W2 R)),
D4,3 = D4,4 = 0. (60)

4. NUMERICAL RESULTS AND DISCUSSIONS


Three different fluid mediums: two liquids (water and glycerin), and one gas (air) are
used in this study. No restrictive assumptions were made regarding the nature of the fluids
when we constituted the fluid governing equations (1)–(5). Their thermal properties [14, 15]
are listed in Table 1. The dimensionless frequency V( = va/C0 , a = 0·0414 m) is in the
range 0·01–12, which corresponds to 57–68 321 Hz in water; 73–87 420 Hz in glycerin; and
13–15 823 Hz in air. We chose these frequency bands because we are more interested in
studying acoustic waves audible by humans.
Three-dimensional mode shapes in terms of velocity of the wave propagating along
z-direction can be expressed as
v(R, u, Z) = Re{v̂(R) einu + zZ}, (61a)
and its corresponding displacement mode shapes can be obtained by dividing it by a
constant iv,
u = v/iv, (61b)
where z = ka(z = zr + izi ) is the dimensionless complex eigenvalue corresponding to a
specific frequency. The constant zr corresponds to the attenuation constant and zi
corresponds to the propagation constant. The dimensionless phase velocity can be
expressed as
p = Cp /C0 = V/ zi .
C* (62)
Table 2 shows the partial results of solving the characteristic equation (57) for z. In
the frequency domain of interest for the fluid system, z is either nearly pure imaginary
which corresponds to a propagation mode, or nearly pure real which corresponds to an
attenuation mode (or evanescent mode). Six propagation modes of the water line are

1110 . .   . . 


presented in the dispersion spectra as Figure 2. Each one has corresponding attenuation
modes except the fundamental mode P10. Three of them P10, P20 and P30, are axially
symmetric modes n = 0. The other three, P11, P21 and P31, are non-axially symmetric
modes n = 1.
A complex mode where the magnitudes of real and imaginary parts of z are in about
the same order, do not appear for the fluid system, since its behavior is different from the
solid system [16] for the same frequency range. The code system used to present each of
the modes in the fluid systems is consistent with the code system used in the solid system
[16]. The mode P10, for example, denotes the propagation mode 1 at n = 0. The mode A31
denotes the attenuation mode 3 at n = 1.
The phase velocity dispersion plots of the water line in Figure 3 show that the phase
velocity of the fundamental mode P10 is essentially independent of frequency, and hence
is non-dispersive with its coinciding with the intrinsic sound speed in water and exists at

Figure 5. Two-dimensional mode shapes of displacement (only the real part is shown) of the water line at
different frequencies for axially symmetric waves, n = 0, and non-axially symmetric waves, n = 1. All waves start
at low frequencies as longitudinal vibrations and for higher frequencies transit into longitudinal-transversal
vibrations except the fundamental mode P10. The fundamental mode is always in longitudinal vibration (plane
wave): - - -, before deformation; ——, after deformation. (a) Mode P10; (b) mode P20; (c) mode P30; (d) mode
P11; (e) mode P21; (f) mode P31.
  1111
all frequencies. All of the higher modes P11, P20, P21, P30 and P31 display minimum
cut-off frequencies at 2·2, 3·833, 5·4, 7·02 and 8·6, respectively. The cut-off frequencies are
dependent on the diameter of the water line. The dispersion spectra and phase velocity
plots for glycerin and air are very close to the plots for water. One cannot tell the difference
graphically between them with the naked eye. However, Table 2 shows the propagation
constant zi of three fluid mediums to be very close. For instance, they are 12·0003, 12·0096
and 12·0025, respectively, at V = 12, mode P10. Also, note that the modes have a very
small attenuation zr .
Figure 4 shows typical one-dimensional mode shapes of displacements, stresses and
temperature in the radial direction for water at V = 4, mode P20 (axially symmetric mode).
The displacements for the r and z components are seen to be zero at the fluid cylindrical
surface which satisfies the no slip boundary conditions. The two-dimensional mode shape
plots of the water line in Figure 5 for axially symmetric waves, n = 0, and for non-axially
symmetric waves, n = 1, show wave vibrations. The nature of the wave vibrations are
dependent on the frequencies. For instance, mode P20 starts as a one-dimensional
longitudinal vibration at V = 3·833, then changes to a two-dimensional, essentially
longitudinal wave with some transverse motion at V = 4 and 12. The vibration of the
fundamental mode, P10, is always one-dimensional longitudinal (plane wave) across the
whole frequency range.
The attenuation rate zr of six propagation modes for water, glycerin and air in Table 2
are distinctly different as a class as is seen in the plots shown in a Figure 6. Generally,
the attenuation rates of the propagating modes are nearly zero. Glycerin has the highest
attenuation rate because it has the largest viscosity. Suprisingly, air does not have the
lowest attenuation rate although its viscosity is the lowest. This is because in addition to
viscosity, the attenuation rate also depends on the intrinsic sound speed and density.
Higher viscosity, slower sound speed and lower density will increase the attenuation rate.
The air viscosity is small but it also has a slower sound speed and lower density than water.
The attenuation rates of glycerin and air are about 37 and 9 times higher than water,
respectively, for all six propagation modes.
The fundamental mode P10 in Figure 6 is the least attenuated mode in the lower
frequency range. When the dimensionless frequency V increases to above 3·84, the second
axially symmetric mode P20 takes over as the least attenuated mode. For air at room
temperature and propagating in 0·0414 m and 1 m radius pipes, V 3·84 corresponds to 5063
and 210 Hz, respectively. The cut-off frequencies of the three fluid mediums are very close.
For instance, for mode P20, the cut-off frequencies of water, glycerin and air are 3·833,
3·875 and 3·85, respectively.
The first non-axially symmetric mode P11 seems never to be the least attenuated mode.

T 3
Non-thermal effect solutions (g0 = 1) at the first two modes P10 and P11 in air line, where
z = (zr ,zi ). The non-thermal effect solutions of air is about identical to thermal effect solutions
in Table 2. (a) Fundamental mode P10; (b) first non-axially symmetric mode P11
Nondimensional Nondimensional
frequency (a) frequency (b)
0·01 (0·000074, 0·01007) 1·85 (0·004270, 0·18473)
1 (0·000728, 1·00073) 2 (0·001510, 0·78255)
4 (0·001468, 4·00145) 4 (0·001991, 3·55303)
8 (0·002116, 8·00205) 8 (0·002945, 7·78813)
12 (0·002654, 12·0025) 12 (0·003691, 11·8615)
1112 . .   . . 

Figure 6. Attenuation rates of propagation modes in water, glycerin and air. Water has the lowest attenuation
rate. The glycerin and air have attenuation rates which are about 37 and 9 times higher than water, respectively:
——, axially symmetric modes, n=0; ...., non-axially symmetric mode, n = 1; —w—, P10; - -w- -, P11; —r—,
P20; - -r- -, P21; —q—, P30; - -q- -, P31.

Once each of the higher non-axially symmetric modes begin to propagate for increasing
frequency, its attenuation is less than its symmetric counterpart. For instance, the
non-axially symmetric mode P21 has a lower attenuation rate than the symmetric mode
P20 in Figure 6.
The attenuation rates with and without thermal effect (if the ratio of specific heats g0
is set to 1 for the non-thermal effect case) do not differ too much. Even though air has
the highest thermal effect because of the highest g0 , the attenuation rate at room
temperature with the thermal effect is only slightly higher (by 2%) than that without the
thermal effect at mode P10, V = 12, and is shown in Table 3. The thermal effect does not
seem to have much influence in water or glycerin. The difference in attenuation rates is
less than 0·001%.

REFERENCES
1. G. K 1868 Annalen der Physik and Chemie 134, 177–193. Uber den Einfluss der
Wärmeleitung in einem Gase auf die Schallbewegung.
2. C. R. G and J. D. P 1967 Transactions of the American Society of Mechanical
Engineers, Journal of Basic Engineering 89, 782–788. Wave propagation in viscous fluid lines
including higher mode effects.
3. H. A. S and W. T. R 1973 Journal of Fluid Mechanics 58 (part 3), 595–621.
Axisymmetric waves in compressible Newtonian liquids contained in rigid tubes: steady-periodic
mode shapes and dispersion by the method of eigenvalleys.
4. H. T 1975 Journal of Sound and Vibration 39, 1–33. On the propoagation of sound waves
in cylindrical tubes.
5. P. H and K. K 1976 Stanford University J.I.A.A. TR-4 . Effects of friction and
heat conduction on sound propagation in ducts. See p. 13.
6. C. Z and C. W. K 1949 Sound Absorbing Materials, New York: Elsevier. See
Chapter 1.
7. P. E. D and P. G. V 1970 Journal of Sound and Vibration 12, 201–224. Attenuation of
plane wave and higher order mode sound propagation in lined ducts.
  1113
8. M. L. M 1987 Acoustics of Ducts and Mufflers. New York: John Wiley. See Chapter 6.6.
9. P. N. L 1993 Noise-Con 93 , 265–272. Effect of mylar surface films on acoustic fill on silencers
performance.
10. P. E. D 1973 Journal of Sound and Vibration 26, 91–120. Analysis of internally generated
sound in continuous materials: 3. The momentum potential field description of fluctuating fluid
motion as a basis for a unified theory of internally generated sound.
11. P. A. M 1979 Journal of Fluid Mechanics 91 (part 2), 357–397. The linearized treatment
of general forced gas oscillations in tubes.
12. C. M. C 1987 Ph.D. Thesis, Rensselaer Polytechnic Institute. Acoustic wave dispersion in
a viscous, compressible, liquid-filled cylindrical elastic tube. (UMI 87-29299)
13. P. N. L 1990 Ph.D. Thesis, Rensselaer Polytechnic Institute. Thermal acoustic wave
propagation within a slightly compressible viscous fluid-filled impermeable cylindrical elastic
tube. See Section 2.3 and 3.5. (UMI 91-11065)
14. E. K. L, A. R. F, A. B. C and J. V. S 1982 Fundamentals of Acoustics.
New York: John Wiley. See pp. 462–463.
15. P. A. T 1972 Compressible-Fluid Dynamics. New York: McGraw-Hill. See p. 644.
16. P. N. L and H. A. S 1994 Journal of Sound and Vibration 177(1), 121–135.
Three-dimensional mode shapes for higher order circumferential thermoelastic waves in an
annular elastic cylinder.

You might also like