Sie sind auf Seite 1von 11

International Applied Mechanics, Vol. 41, No.

10, 2005

ON FREE NONLINEAR VIBRATIONS OF FLUID-FILLED CYLINDRICAL SHELLS


WITH MULTIPLE NATURAL FREQUENCIES

UDC 539.3

V. D. Kubenko, P. S. Kovalchuk, and L. A. Kruk

Nonlinear free vibrations of a cylindrical shell fully filled with a perfect incompressible fluid are studied.
The case is examined where two natural frequencies of the shell are close
Keywords: cylindrical shell, perfect incompressible fluid, nonlinear vibrations, first integrals, traveling wave
Introduction. The nonlinear vibrations of fluid-filled thin shells represented by simple (usually one-degree-of-freedom)
models have been studied in sufficient detail (see [1, 2, 7, 9, 12, etc.] for review of such studies). However, such models fail to
describe many complex nonlinear phenomena associated with the interaction of two or more different modes of shells filled with
a fluid. The coupling of such modes may be a cause of intensive energy exchange among them, in addition to specific
(resonant) conditions. This may result in qualitatively new deformation modes differed from traditional standing waves. These
new modes are, in particular, waves traveling in the circumferential direction. We also investigate a superposition of standing
and traveling waves, chaotic modes, and other phenomena.
The present paper sets forth a method for analysis of the free multimode vibrations with internal resonances of
finite-length cylindrical shells filled with a fluid.
1. Initial Dynamic Equations. For dynamic equations of a fluid-filled cylindrical shell we use mixed geometrically
nonlinear equations of the theory of flexible shallow shells [14]:
2
2
2
2
2
2
2
D 4
&& = W + W 2 W + 1 Ph ( x, y, t ) ,
W + W
h
xy xy R x 2
h
x 2 y 2 y 2 x 2
2

2W 2W 2W 1 2W
1 4
2
,

=
E
x y 2 R x 2
xy

(1.1)

where W = W(x, y, t) is the radial deflection of the shell (positive when directed inward); x and y are the longitudinal and
circumferential coordinates, respectively; the origin of the X-axis is at the center of one of the shell ends; R is the mean radius of
the shell; D = Eh3/(12(1 2)) is the flexural rigidity of the shell (E is Youngs modulus, h is the shell thickness, and is
Poissons ratio); is the mass density of the shell; = (x, y, t) is the in-plane stress function; Ph is the internal fluid pressure
applied to the lateral surfaces of the shell; 4 is a biharmonic operator defined as 4 = (2/x2 + 2/y2)2; and W& = W/t,
&& = 2W/t2.
W
We assume that the shell is simply supported at the ends (SS1/SS1):
W = 0,

M x = 0,

N x = 0,

V = 0 at

x = 0, l,

(1.2)

where Mx is the bending moment per unit length, Nx is the force per unit length in the axial direction, V is the displacement of the
mid-surface in the circumferential direction; and l is the length of the shell. The prestress of the shell due to the weight of the fluid
is neglected [2].

S. P. Timoshenko Institute of Mechanics, National Academy of Sciences of Ukraine, Kiev. Published in Prikladnaya
Mekhanika, Vol. 41, No. 10, pp. 127138, October 2005. Original article submitted October 22, 2004.
1063-7095/05/4110-1193 2005 Springer Science+Business Media, Inc.

1193

Since the shell is closed in the circumferential direction, its dynamic flexural deformation can be represented as a
two-parameter expansion:

W=

[ f1n ,m ( t )cos sn y + f 2n ,m (t )sin sn y] sin m x,

(1.3)

n= 0 m=1

n ,m

where sn = n/R and m = m/l are parameters describing the wave shape, and f i

( t ) are unknown time-dependent functions,

which represent generalized coordinates.


The wave representation (1.3) is convenient for analyzing the resonant nonlinear vibrations of the shell [5, 13]:

W=

a nm ( t )cos( sn y nm ( t ))sin m x,

(1.4)

n= 0 m=1

where
a nm ( t ) =

( f1nm (t )) 2 + ( f 2nm (t )) 2 ,

nm ( t ) = arctan

f 2nm ( t )
f1nm ( t )

(1.5)

are the amplitude and phase of vibrations propagating in the circumferential direction.
To determine the internal pressure on the shell, we use Bernoullis equation from which it follows that
Ph = 0

( x, r, , t )
,
r= R
t

(1.6)

where 0 is the mass density of the fluid; x, r, and are the cylindrical coordinates; and y = R.
The velocity potential is determined from the boundary-value problem [2, 9, 12]
2 2 1 1 2
+
+
+
= 0 in
x 2 r 2 r r r 2 2

r= 0

< ,

r= R

W
,
t

x= 0 = 0,

Q,
x= l = 0,

(1.7)

where Q is the domain occupied by the fluid.


Solving the boundary-value problem (1.7) and taking the expansion (1.4) into account, we get the following expression
for the unknown function Ph:

Ph = 0

[M nm cos( sn y nm )+ N nm sin(sn y nm )]

n= 0 m=1

I n ( m r )
sin m x,
m I n ( m r )

(1.8)

where
M nm = a&& nm a nm & 2nm ,

&& nm + 2a& nm & nm ,


N nm = a nm

(1.9)

In(mr) is a modified Bessel function, and I n ( m r ) is the derivative of In(mr).


We substitute the expansion of W (1.4) into the second equation in (1.1) and then determine the function , which, in
view of (1.2), can be represented as follows:
= h + p ,
where h = Kx2/2 is a homogeneous solution and p is a particular solution:

1194

(1.10)

cos( sn y nm )sin m x +

p =
nm
0
n= 0 m=1

nmjk

nmjk

{ 1nmjk [cos(sn s j ) y ( nm jk )]

n= 0 m=1 j= 0 k=1

cos( m k )x + 2
+ 3

[cos(sn + s j ) y ( nm + jk )] cos( m k ) x

[cos(sn s j ) y ( nm jk )] cos( m + k )x + nmjk


[cos(sn + s j ) y ( nm + jk )]cos( m + k )x}.
4
nmjk

The functions nm
0 and 0

nm
0 =

nmjk

(1.11)

(i = 1, ..., 4) are defined by


E2m a nm
,
R ( m , sn )

nmjk

E ( k sn + m s j ) 2 a nm a jk
4 ( m k , sn + s j )
nmjk

E ( k sn m s j ) 2 a nm a jk
4 ( m k , sn s j )
nmjk

E ( k sn + m s j ) 2 a nm a jk
4 ( m + k , sn s j )

E ( k sn m s j ) 2 a nm a jk
4 ( m + k , sn + s j )

(1.12)

and the operator is defined by (A, B) = (A2 + B2)2.

2R

The unknown function K appearing in (1.10) can be determined from the periodicity condition

V
dy = 0 [14]:
y

E
K = sn2 a nm .
8 n= 0 m=1

(1.13)

Using the BubnovGalerkin procedure and choosing cos(sny nm)sinmx and sin(sny nm)sinmx as weighting
functions, we obtain the following system of coupled nonlinear equations:
nm
2 )a
a&& nm + ( 2nm & nm
nm = F1 ,

&& nm + 2a& nm & nm = F2nm


a nm
(0 )2

(n = 0, 1, 2, ..., m = 1, 2, ...),

(1.14)

(0 )

where 2nm = nm / (1+ M 0nm ), nm is the vibration frequency of the shell without fluid,
(0 )
nm


E4m
1 D
= (2m + sn2 ) 2 + 2 2

2
2
R ( m + sn )
h

1/ 2

(1.15)

M 0nm is the added mass parameter of the fluid


M 0nm =

0 I n ( m R )
,
m hI n ( m R )

(1.16)

Finm (i = 1, 2) are power functions (up to the third order inclusive) of the variables {apq} = {a01, a02, ...}, which are due to the
nonlinear terms in the equations [13].
The formal parameter (0 < << 1) introduced in (1.14) shows how small the functions Finm are compared with the
right-hand sides of these equations. As shown in [4, 13], the parameter can be represented as = Wmax/R (Wmax is the maximum
radial deflection of the shell) because all the nonlinear terms in (1.1) are proportional to it.

1195

Analyzing the system of equations (1.14), we can reveal all main laws governing the free multimode vibrations of
fluid-filled shells with large amplitudes. We can also study special features of the nonlinear modal interaction at the frequency of
internal (multiple) resonance.
2. Periodic Solutions in the Case of Internal Resonances. Let the condition n1 m1 ( p / q) n2 m 2 (where p and q are
mutually distinct numbers, n1 n2) be satisfied for the frequencies nm. To construct approximate periodic solutions of the
system (1.14), we will apply the asymptotic averaging method developed by Bogolyubov and Mitropolsky [3]. We will also use
the results we obtained and published in [5, 6, 811]. The solutions of Eqs. (1.14) will be constructed separately for resonating
and nonresonating waves (with indices n1, m1 and n2, m2) in the approximation (1.4). In the former case, we change variables as
follows:
a& nm =

a nm = A nm + B nm sin nm ,

nm = nm + arctan

B nm
cos nm ,
a nm

nm
+ B nm
2
,
Tnm

A nm tan

& nm =

Tnm
2nm

(2.1)

2 B 2 (A B 0); = ; and A , B , ,
where n = n1, m = m1 and n = n2, m = m2; nm = 2(t + nm ); Tnm = A nm
nm
nm
nm
nm nm nm
nm

and nm are new time-dependent variables to be determined. Determining them as in [8, 9], we obtain the following system of
equations in standard form [3]:
1
dA nm
=
dt
a nm
1
dB nm
=
dt
a nm

d nm
1
=
dt
2a nm B nm
1
d nm
=
dt
2a nm

(1)
Fnm (K )Tnm

[Fnm(1) (K )B nm cos nm + Fnm(2) (K )Tnm ] = U nm ,

[Fnm(1) (K )A nm cos nm Fnm(2) (K )Tnm sin nm ] = Vnm ,


[Fnm(1) (K )( A nm sin nm + B nm )+ Fnm(2) (K )Tnm cos nm ] = nm(1) ,

sin nm cos nm

A nm
B nm

A nm
B nm
(2 )
+ Fnm
(K ) sin nm +
cos nm +
B nm
A nm


(2 )
= nm
. (2.2)

The first four equations result when n = n1 and m = m1, and the other equations when n = n2 and m = m2 and the functions
(1)
Fnm

() are replaced by Fn2 m 2 (... )+ a n2 m 2 , where = 2n

1 m1

n2

2m2

How to construct asymptotic solutions of equations similar to (2.2), which can be written in vector form as dx/dt =
X(x, t), was explained in detail by Bogolyubov and Mitropolsky [3]. According to their method, the right-hand sides of these
equations are represented as sums of harmonic components:

U nm = U nmj {A pq },{B pq } cos j nm +U nmj {A pq },{B pq } sin j nm


(1)

(2 )

(p = 0, 1, 2, ; q = 1, 2, ),

(2.3)

and so on.
In the first approximation, the solutions of Eqs. (2.2) are given by
(1)

A nm = A nm ,
and
1196

(1)

B nm = B nm ,

(1)

nm = nm ,

(1)

nm = nm ,

(2.4)

(1)

(1)

dA nm
(1)
(1)
(1)
= U nm = U nm 0 {A pq },{B pq } ,
dt

(1)

(1)

d nm
(1)
= nm ,
dt

dB nm
= Vnm ,
dt

d nm
(2 )
= nm ,
dt

(2.5)

where the overbar denotes averaging over time t for the functions explicitly appearing in the expression.
To obtain the second approximation, it is necessary to add the small vibrational terms entering into the expressions for
Anm, Bnm, nm, and nm to the first approximation:

A nm = A nm + U nmj {A pq },{B pq } sin j nm U nmj {A pq },{B pq } cos j nm ,


(1)

(1)

j= 0

(1)

(1)

(1)

(1)

nm = 2( nm t + nm )

(1)

(2 )

(p = 0, 1, ...,

(1)

(1)

(1)

q = 1, 2, ...),

(2.6)

and so on.
Substituting (2.6) into the right-hand side of Eqs. (2.2) and averaging the result over time t, we determine Anm, Bnm,
nm, and nm up to small quantities of the second order.
A similar method can be employed to construct asymptotic solutions to the simultaneous equations (1.14) when no
internal resonance occurs, i.e., when two or more natural frequencies nm (n = 0, 1, 2, ; m = 1, 2, ) of the shell are neither
adjacent nor multiple. To this end, we put = nm in (2.1) and (2.5).
3. Two-Wave Model of Shell. To explain the method, we will analyze the vibrations of a fluid-filled shell modeled by a
system with five degrees of freedom. The dynamic deflection W is approximated by the expansion
W = a1 cos( s1 y 1 )sin m x + a 2 cos( s2 y 2 )sin m x + c sin 4 m x
(si = ni/R,

i = 1, 2,

n1 n2,

m = m/l),

(3.1)

which is derived by taking the first four terms in (1.3) and adding the correction term csin4mx. Some authors [14, 13, etc.] used
the function sin2mx to correct the deflection W of a simply supported shell. However, this function only partially satisfies the
boundary conditions (1.2) (specifically, the condition Mx = 0 does not hold). The axisymmetric deflection component sin4mx in
(3.1) is the simplest function that satisfies all the conditions (1.2) and, at the same time, describes the damping of vibrations with
large deflections. The effect of preferential inward buckling observed in this case is because the length of the middle
cross-section outline practically does not change during vibrations with large deflections.
In view of (3.1), the system of equations (1.14) is written as
a&&1 + ( 12 & 12 )a1 + k11 a13 + k12 a1 a 22 + k13 a1 c + k14 a1 c 2 = 0,
a&& 2 + ( 22 & 22 )a 2 + k 21 a 2 a12 + k 22 a 23 + k 23 a 2 c + k 24 a 2 c 2 = 0,
&& 1 + 2a& 1& 1 = 0,
a1

&& 2 + 2a& 2 & 2 = 0,


a2

&&c + 23 c + k 31 a12 + k 32 a 22 + k 33 ca12 + k 34 ca 22 = 0,

(3.2)

where i (i = 1, 2, 3) are the natural frequencies of the shell with fluid,


2j =

1
m0 j

23 =

D
E4m
( m , s j ) + 2
R ( m , s j
h

64 8D4m 35 E
+
,

64 R 2
35m03 h

,
)

m03 = 1+

m0 j = 1+

0 I n j ( m R )
m hI n j ( m R )

16 0 1
35 hl 2

k =1

(j = 1, 2)

I 0 ( k R )M k2 (1 (1) k )
k I 0 ( k R )

1197

Mk =

1924m
k (2k

42m )(2k

162m )

k =

k
,
l

(3.3)

kpq (p = 1, 2, 3; q = 1, 2, 3, 4) are constant coefficients


k11 =

E
(4 + 3s14 ),
16m01
k14 =

k 23 =

m01

Es22
Rm02

k 32 =

k 01 =

E4 s14 k 02

k12 =

E
( s 2 s 2 + k 01 ),
8m01 1 2

k 21 =

E
( s 2 s 2 + k 01 ),
8m02 1 2

5
24
+
,
8 ( , s2 )

2
16 Es2
35 Rm03

k 24 =

5
24
+
,
8 ( , s2 )

E4 s24 k 03
m02
k 33 =

k13 =

k 22 =

k 31 =

Es12 5
24
+
,
Rm01 8 ( , s1 )

E
(4 + 3s24 ),
16m02
2
16 Es1
35 Rm03

4
32 E4 s1 k 02
,
35 m03

k 34 =

5
24
+
,
8 ( , s1 )

4
32 E4 s2 k 03
,
35 m03

1
1
1
1
4
+
+
+ ( s1 + s2 ) 4
( s s )4
,
2 1 2 ( 2 , s1 + s2 ) ( 0, s1 s2 )
s
s
s
s

(
0
,
+
)

(
2
,

1
2
1
2

k 02 =

1
4
1
,
+
+
( , s1 ) ( 3 , s1 ) ( 5 , s1 )

k 03 =

1
4
1
,
+
+
( , s2 ) ( 3 , s2 ) ( 5 , s2 )

= m.

(3.4)

According to [14], we can determine the function c = c(t) from the quasistatic variant of the problem. Neglecting the
inertial term in the last equation in (3.2), we obtain the relation
c=

k 31 a12 + k 32 a 22
23 + k 33 a12 + k 34 a 22

(3.5)

Substituting it into the first two equations in (3.2) and retaining terms up to the third order in a1 and a2, we obtain the
system of equations
a&&1 + ( 12 & 12 )a1 + 1 a13 + 1 a1 a 22 = 0,
&& 1 + 2a& 1& 1 = 0,
a1

a&& 2 + ( 22 & 22 )a 2 + 2 a 23 + 2 a 2 a12 = 0,


&& 2 + 2a& 2 & 2 = 0.
a2

(3.6)

It describes the interaction of natural bending modes of the fluid-filled shell. Here
i =

i =

E 2 si4
16
E
(4 + 3si4 )
16m0 i
35 23 Rm0 i m03

E 2 s12 s22
16
E
( s12 s22 + k 01 )
8m0 i
35 23 Rm0 i m03

5
24
+
8 ( , si

,
)

5
24 5
24
+
+
,
8 ( , s1 ) 8 ( , s2 )

(3.7)

where i = 1, 2.
We can integrate the last two equations in (3.6) independently of the first two equations. As a result we deduce the
integrals
a12 & 1 = C1 ,
where C1 and C2 are the constants of integration,
1198

a 22 & 2 = C 2 ,

(3.8)

C1 = a12 ( 0)& 1 ( 0),

C 2 = a 22 ( 0)& 2 ( 0).

(3.9)

Thus, the problem of free nonlinear vibrations of a fluid-filled shell with deflections approximated by (3.1) is reduced to
the following system of equations:
a&&1 + 12 a1 + 1 a13 + 1 a1 a 22 = C1 / a13 ,

a&& 2 + 22 a 2 + 2 a 23 + 2 a 2 a12 = C 2 / a 23 .

(3.10)

The general form of shell deformation is strongly dependent on the initial conditions & i ( 0) (i = 1, 2) and the relationship
between the frequencies 1 and 2.
4. Resonant Vibrations of Shell at 1 2. Analyzing the system of equations (3.10) in a certain way [3], we conclude
that the first approximation reveals only the principal internal resonance 1 2. The approximate periodic solution (2.1) for
this system may be represented as
a i = A i + B i sin i ,

i = 2(t + i ),

= 1

(i = 1, 2),

(4.1)

where Ai and Bi are time-dependent functions such that


A i2 B i2 =

C i2
2

(4.2)

The unknown parameters Ai and i can be determined from the following system of averaged equations:
1
dA1
=
B B sin ,
2 1 1 2
dt

dA 2 2
=
B B sin ,
2 1 1 2
dt

d 1
B2

1
3 1 A1 + 2 1 A 2 + 1 A1
=
cos ,
dt
B1
4 1

d 2
B1

1
3 A + 2 2 A1 + 2 + 2 A 2
cos ,
=
dt
B2
4 1 2 2

(4.3)

where = 2 1 and = 22 12 .
Then it is easy to obtain the first integral from the first two equations of (4.3):
A1 A 2
+
= C 0 = const,
1 2

(4.4)

which describes the energy relationship, due to vibrations, of the amplitude parameters A1 and A2. Namely, as one of these
parameters increases, the other decreases, and vice versa. The constant of integration C0 is given by
C0 =

1 1 2
1
[ a& 1 ( 0) + a12 ( 0)( 12 + & 12 ( 0)] + [ a& 22 ( 0) + a 22 ( 0)( 12 + & 22 ( 0)].
2
2

1
1

(4.5)

Using this integral, we reduce Eqs. (4.3) to


1
dA`1
=
L ( A )L ( A )sin ,
2 1 1 `1 2 `1
dt

1
d
= N 0 + N 1 A1
S ( A1 )cos ,
2 1
dt

(4.6)

where
2

A1

L1 =
C 0 22 C 2 ,
1

L2 = A12 C1 ,

S=

1
L1 L2

2 2 2
A1 L1 L2 C 0 2 A1 ,
1
1

1199

TABLE 1
n

0i , Hz

779.7

476.8

335.7

283.8

291.4

337.1

406.2

i , Hz

209.7

144.7

111.99

102.2

111.96

136.8

172.9

TABLE 2

N0 =

l/R

2.45

2.40

2.50

03 , Hz

5037.0

5037.0

5037.0

3 , Hz

925.18

934.44

916.9

1
[( 3 2 21 ) C 0 2 + 2 ],
2

N1 =

2
1
4 2 ,
3 1 + 2
1
2

1 = 12 / 2 ,

2 = 22 / 2 .

(4.7)

From here we deduce a more general integral:


N 0 A1 +

N 1 A12
2

1
L L cos = C 3 = const,
2 1 2

(4.8)

which relates the amplitude and phase parameters of the shellfluid system.
In general, the integrals (4.4) and (4.8) enable us to trace the redistribution of energy, given at some initial time, between
the bending modes, which is because of nonlinear relationships and internal resonances. Evidently, the amplitude A1 = A1(t) and
the phase 1 = 1(t) are determinative parameters. They make it possible to fully study the processes of modal interaction.
Indeed, if we know the dependence A1 = A1(t), then we can find from (4.2) and (4.4) the values of A2(t), B1(t), and B2(t). And
from the integral (4.8) we can determine the values of the functions (t) and 2(t).
Since the amplitude and phase parameters are now known, we can find the total amplitudes ai and phases i from (2.1)
and the amplitude c from the general vibration process described by (3.1). When the values of ai are known, it is easy to
determine the phase velocities vi (i = 1, 2) of traveling waves with the parameters (s1, ) and (s2, ):
v1 =

& 1 T1 1
=
,
s1 a 2 s1
1

v1 =

& 2 T2 1
.
=
s2 a 2 s2
1

(4.9)

If there is no fluid, then the integrals (4.4) and (4.8) can be simplified, since 1 = 2 = in this case.
If the resonance condition 1 2 does not hold, then the modes with wave parameters m, n1 and m, n2 do not interact
intensively. In this case, Eqs. (3.1) are not coupled because 1 = 2 = 0. To study the interaction of the modes at 1 2, it is
necessary to take into account the second and higher approximations [10].
5. Numerical Results. Let a fluid-filled (0 = 103 kg/m3) shell be characterized by the following parameters:
E = 2 1011 Pa,

= 7.8 103 kg/m3,

h / R = 3.125 103 ,

l / R = 2.45,

R = 0.16 m,

= 0.3.

(5.1)

Table 1 contains the natural frequencies i = i/2 of this shell for m = 1 and different values of n. For comparison, the
table includes (in the first row) the frequencies 0i = 0i/2 of the same, yet dry shell.

1200

A1

d / dt ,
rad/sec

0.9

0.9

0.7

0.7

0.5

0.5
0

0.05

0.10

0.15 t, sec

0.05

0.10

0.15 t, sec

b
Fig. 1
d / dt ,
rad/sec

A1
6

1500

5
1000
4
500

2
1

1
500
0

0.02

0.04

t, sec

0.02

0.04

t, sec

b
Fig. 2

Table 2 collects the frequencies 3 corresponding to the axisymmetric mode sin4mx of the fluid-filled shell for
different values of l/R.
The frequency 03 = 03/2 corresponds to the dry shell with m03 = 1.
As demonstrated by Table 1, the shell with fluid has two, almost equal frequencies (1 2) for modes with the wave
parameters m = 1, n1 = 5 and m = 1, n2 = 7. The frequency 3 = 3/2 (Table 2) is much higher than the frequencies i (i = 1; 2)
for the parameters under consideration. The other frequencies of the empty or fluid-filled shell do not satisfy the
internal-resonance conditions.
Figure 1 presents the results from the numerical integration of the first two equations in (4.6) by the fourth-order
RungeKutta method (with integration step chosen automatically) with the parameters (5.1), n1 = 5, n2 = 7, and the initial
conditions & 1 ( 0) = & 2 ( 0) = 0, a1(0) = a2(0) = h, and ( 0) = / 4.
Here A1 is the dimensionless amplitude parameter (A1 = A1/h2 because the dimensionless amplitude is
a1 = a1 / h = ( A1 + B1 cos 1 ) / h 2 ). Note that with the initial conditions & i ( 0) = 0 (i = 1, 2), the deformation of the shell is the
superposition of two standing waves with the amplitudes a1 and a2 (which follows from (3.1) and (3.8)).
Figure 2 illustrates the influence of the initial amplitudes a1 and a2 on the energy exchange between the mode m = 1,
n1 = 5 and the mode m = 1, n1 = 7.
The functions A1 ( t ) and & ( t ) plotted for the dry (Fig. 1a) and fluid-filled (Fig.1b) shells with nonzero initial conditions
(0 )
for & i ( 0), i.e., & 1 ( 0) = & 2 ( 0) = nm and a1(0) = a2(0) = h, ( 0) = / 4. With such initial conditions, the deformation process is
the superposition and interaction of two traveling bending waves (since & 1 ( t ) 0 and & 2 ( t ) 0 when t > 0).

1201

dA1 / dt

dA1 / dt

6.4
3.2
3.2

1
2

3.2

3.2
6.4
9.6
0.032 0.064 0.096
a
dA1 / dt

6.4
0.064 0.128
A1
b
dA1 / dt
3

0.192

A1

25.6
128

12.8
2

0
2

12.8
128
25.6
38.4

256
0.128

0.256

A1

0.512

0.768

A1

d
Fig. 3

The curves in Figs. 1 and 2 demonstrate that energy exchange between the modes is a periodic process. Its distinguished
feature is that an increase in the amplitude A1 leads to a decrease in the phase velocity d/dt, and vice versa. The fluid changes
& Their oscillations are essentially different from sinusoidal, as
qualitatively the oscillatory behavior of the parameters A1 and .
shown in Figs. 1b and 2b, for instance.
Figure 3 shows the phase trajectories for the periodic process of energy exchange between different modes. These
trajectories are obtained from the integration of Eqs. (4.5) with initial conditions of two kinds: & 1 ( 0) = & 2 ( 0) = 0 (Fig. 3a, b) and
& 1 ( 0) = & 2 ( 0) = (Fig. 3c, d). The phase portraits shown in Fig. 3a, c describe the free vibrations of the dry shell. The phase
portraits in Fig. 3b, d represent the fluid-filled shell. Curves 1 correspond to n1 m = n2 m , l/R = 2.45 (resonance), and curves 2
and 3 to l/R = 2.5 and l/R = 2.4, respectively. The natural frequencies of the fluid-filled shell are somewhat different (for example,
n1 m = 108.33 Hz and n2 m = 110.49 Hz for l/R = 2.5; n1 m = 115.88 Hz and n2 m = 113.54 Hz for l/R = 2.45, where n1 = 5 and
n2 = 7).
Thus, as is seen from Fig. 3, the phase trajectories of shell points during free vibrations are qualitatively different in the
cases & i ( t ) = 0and & i ( t ) 0(i = 1, 2), i.e., when two standing waves or two waves traveling in the circumferential direction are
initially excited. The shape and position of the trajectories on the phase plane strongly depend on the disturbance of the
frequencies 1 and 2 due to some change in the shell geometry (specifically, its length l) in the case being considered. The
trajectory corresponding to zero disturbance (1 = 2) is the so-called watershed separating the trajectories corresponding to
(1 2) > 0 and (1 2) < 0.

REFERENCES
1. M. Amabili and M. P. Padoussis, Review of studies on geometrically nonlinear vibrations and dynamics of circular
cylindrical shells and panels, with and without fluidstructure interaction, Appl. Mech. Rev., 56, No. 4, 349381 (2003).
2. M. Amabili, F. Pellicano, and A. F. Vakakis, Nonlinear vibrations and multiple resonances of fluid-filled circular
cylindrical shells. Part 1: Equations of motion and numerical results, J. Vibr. Acoust., 122, 346354 (2000).
1202

3. 1. N. N. Bogolyubov and Y. A. Mitropolsky, Asymptotic Methods in the Theory of Nonlinear Oscillations, Gordon and
Breach, New York (1962).
4. I. C. Chen and C. D. Babcock, Nonlinear vibration of cylindrical shells, AIAA J., 13, No. 7, 868876 (1975).
5. P. S. Kovalchuk and V. G. Filin, Circumferential traveling waves in filled cylindrical shells, Int. Appl. Mech., 39,
No. 2, 192196 (2003).
6. P. S. Kovalchuk, N. P. Podchasov, and V. V. Kholopova, Periodic modes in the forced nonlinear vibrations of filled
cylindrical shells with an initial deflection, Int. Appl. Mech., 38, No. 6, 716722 (2002).
7. V. D. Kubenko and P. S. Kovalchuk, Nonlinear problems of the dynamics of elastic shells partially filled with a
liquid, Int. Appl. Mech., 36, No. 4, 421448 (2000).
8. V. D. Kubenko and P. S. Kovalchuk, Application of the method of averaging for investigation of nonlinear waves
processes in elastic systems with rotational symmetry, Ukrainian Math. J., 53, No. 10, 13581367 (2001).
9. V. D. Kubenko, P. S. Kovalchuk, L. G. Boyarshyna, et al., Nonlinear Dynamics of Axisymmetric Bodies Containing
Liquid [in Russian], Naukova Dumka, Kiev (1992).
10. V. D. Kubenko, P. S. Kovalchuk, and L. A. Kruk, On multimode nonlinear vibrations of filled cylindrical shells, Int.
Appl. Mech., 39, No. 1, 8592 (2003).
11. V. D. Kubenko, P. S. Kovalchuk, and L. A. Kruk, Non-linear interaction of bending deformations of free-oscillating
cylindrical shells, J. Sound Vibr., No. 265, 245268 (2003).
12. V. D. Kubenko, V. D. Lakiza, V. S. Pavlovskii, and N. A. Pelykh, Dynamics on Elastic GasLiquid Systems under
Vibratory Actions [in Russian], Naukova Dumka, Kiev (1989).
13. V. D. Kubenko, P. S. Kovalchuk, and N. P. Podchasov, Nonlinear Vibrations of Cylindrical Shells [in Russian],
Vyshcha Shkola, Kiev (1989).
14. A. S. Volmir, Nonlinear Dynamics of Plates and Shells [in Russian], Nauka, Moscow (1972).

1203

Das könnte Ihnen auch gefallen