Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Journal of Sound and Vibration (1988) 127(l), 133-143

NON-LINEAR VIBRATION ANALYSIS OF FLUID-FILLED


CYLINDRICAL SHELLS

P. B. GONCALVES

Department of Civil Engineering, Pontificia Universidade Catblica-PUC/ RJ, Rio de Janeiro 22453,
Brasil

AND

R. C. BATISTA

Department of Civil Engineering, COPPEI Universidade Federal do Rio de Janeiro, Rio de Janeiro,
CP 68506-21945, Brasil

(Received 15 March 1988, and in revised form 27 May 1988)

A theoretical analysis is presented for determining the elastic non-linear vibrations


of a prestressed thin-walled cylindrical shell filled with an ideal fluid. For the vibrations
of the shell itself, the dynamic version of the Sanders non-linear equations for the case
of moderately small rotations is employed. Modal expansions are used for the displace-
ments of the shell middle surface that are required to satisfy the “classical simply
supported” boundary conditions and the circumferential periodicity condition. The fluid
is taken as non-viscous and incompressible, and the coupling between the deformable
shell and this medium is taken into account. The velocity potential is expanded in terms
of harmonic functions which satisfy the Laplace equation term by term. The Galerkin
method is used to reduce the problem to a system of coupled algebraic non-linear equations
for the modal amplitudes. Solutions are presented to show the effects of fluid and shell
parameters on the non-linear vibrations of the shell.

1. INTRODUCTION
Thin cylindrical shells are often used as containment vessels or tanks for the storage of
fluids. Prediction of the dynamic behaviour of these fluid containers requires the simul-
taneous consideration of the governing equations of the shell and fluid, and the boundary
conditions at the shell-fluid interface.
Since the early efforts of Rayleigh and Love, the linear vibrations of shells in a vacuum
have been extensively analyzed [ 1,2]. The linear dynamic interaction between shells and
fluid has also received a good deal of attention [3-51.
Recently, due to the growing appreciation of the importance of non-linear effects upon
the stability and response of thin shells under dynamic loading, a great amount of effort
has been put forth to investigate the non-linear vibrations of thin shells. Although
numerous papers have been written on the non-linear vibrations of cylindrical shells in
air, on the basis of various shell theories [6-111, no such non-linear analysis appears to
have been consistently carried out for fluid-filled cylindrical shells.
The purpose of the present work is to present a mathematical model for the non-linear
dynamic interaction between a fluid and a thin elastic shell, and to investigate the effects
of shell and fluid parameters on the non-linear structural vibration response.
In the present investigation the non-linear dynamic version of Sanders’ equations for
small strains and moderately small rotations is used [12]. Because of the inclusion of
133
0022-460X/88/220133+ 11 %03.00/O @ 1988 Academic Press Limited
134 P. H. CONCALVES AND R. (‘. BATISTA

in-plane inertia effects and the avoidance of the approximations of the shallow shell
theory [13], generally used in non-linear analyses of shells, the results are applicable to
a wide range of wavelengths and also for cases of excitation at the higher frequencies of
the shell. The middle surface displacements are expanded in terms of harmonic functions
which are required to satisfy the classical simply supported boundary conditions and
continuity requirements exactly.
The fluid is treated as non-viscous and incompressible and its motion is assumed to
be irrotational. As a result, it can be characterized by a velocity potential. The solution
for the velocity potential is taken as a sum of suitable functions, the unknown parameters
of which are determined by the kinetic condition along the wetted surface of the shell.
Possible fluid non-linearities will generally come into play when the the vibration
amplitudes are significant. In reference [14] the non-linear terms in the dynamic fluid
pressure and in the boundary conditions at the surface of the cylinder were considered.
That study, which explored many aspects of the cylindrical shell-fluid interaction, demon-
strated that these effects are of secondary order for completely filled or submerged shells,
provided that the movement of the shell surface is moderately small (up to the order of
a few times the shell thickness). Thus, the non-linear terms in the fluid equations have
been neglected in the present investigation. The negligible effect of the non-linear terms
in the boundary condition at the solid-fluid interface was also inferred by Ginsberg when
studying the propagation of weakly non-linear waves [ 15,161.

2. PROBLEM FORMULATION
This section describes the equations to model the shell and the fluid region inside the
shell. The cylindrical co-ordinates (x, 0, r), geometrical parameters (L, R, h) and positive
directions of the displacement components (U, V, W) used in the following analysis are
shown in Figure 1.

Figure 1. Co-ordinate system, displacements and geometry.

2.1. SHELL EQUATIONS

It is assumed that an isotropic, elastic, thin-walled cylindrical shell segment,,filled with


an ideal fluid, initially at rest, is subjected to a static axial load PO and a constant lateral
pressure po. The displacement field associated with the vibrations of this prestressed shell
consists of an axisymmetric static prestress displacement field U. = { U,, V,, W,}, which
occurs prior to excitation, plus an additional field U = { U, V, W} resulting from the
excitation.
NON-LINEAR FLUID-SHELL INTERACTION 135

For the sake of simplicity, the bending confined to narrow zones adjacent to the supports
is neglected, and the prestress displacement field of the shell is taken as a linear membrane
state. In dimensionless form, it may be expressed as
U3=(n,cB- w3o)[5--(Ka1/2~, v,=o, w, = (n&)-vn,J/26. (1)
In the foregoing, the following non-dimensional parameters have been introduced:
S = h/2R, 1= L/R, 5=x/R,
n,, = Po( 1 - v2)/2rREh, n,, = poR( 1 - v*)/Eh. (2)
Here E is the modulus of elasticity and v is the Poisson ratio.
For the motion of this prestressed shell, the dynamic version of Sanders’ non-linear
equations for the case of small strains and moderately small rotations is to be used, with
only the rotation about the normal in the expressions for the middle surface strains
neglected [ 121. These differential equations of motion relative to the incremental displace-
ments’ components can be written in the form

-3Aa[~,~,+ (u.0 - u,&/4l.s - A[(w,, - YO)+ ~w,& - G,(W,O- u)


-y* a2v/at2+p0 = 0, (3b)
~,~+u,~+W+S[(w~~-u)2+vw~*]-{26w.*[u,,+Sw:*+v(u,,+w)+S~(w.~-u)zl~,~
--{2s(w,,-u)[(y,+w)++y-d2+ qg+~~wf*ll,e
-c=J(w,,- u)[u,e+ y*+2~w,,(w,,- U)1),6
-Wa~,~[u,,+ u,,+~~w,,(w., - ~)I).B+ NW,,+ ~W,BB- YO)I,N
+A[(w,cw - ye) + ~~,&e+4Mw,,e+ (u,, -3~s)/4l,se- nxow,g- GO(WOB
- use)
+~ZaZ~/at2-(~,++BF)=0. (3c)
Here u, u, w are, respectively, the axial, circumferential and radial displacements of
the middle surface of the shell, non-dimensionalized with respect to the shell thickness,
h, A = h2/12R2, a = (1- v)/2, t is time and y is a time parameter defined by y =
[psR2(l - v’)/E]“*, where ps is the shell material density.
The dimensionless surface loads are expressed by A, &, and h in their positive
directions. The additional pressure pF in the third equation is the dynamic fluid pressure.
These dimensionless components of surface loads are converted to dimensional form by
~i=[4ES~/(l-~~)]pi. (4)
In this investigation, the shell boundary conditions are assumed to be the “classical
simply supported” boundary conditions; thus
U.6= u = w = w,ct = 0 at 5 = 0,l. (5)
In order to prevent rigid body motion, the following requirement must be added:
u=O at 5 = l/2. (6)
136 P. E. CONCALVE. AND R. C. BATISTA

In addition, for a complete cylindrical shell, as is considered here, all components of


displacements must satisfy the circumferential periodicity condition. This condition
assures that all the physical quantities will be continuous and single-valued along the
circumferential co-ordinate.

2.2. FLUID EQUATIONS

irrotational
The motion of an incompressible and non-viscous fluid can be described
by a velocity potential 4(x, r, 0, t) [ 17,181. This potential function must satisfy the Laplace
equation which may be written in dimensionless form as

~,~~+(~.J~)+(~,,ee/~2)+~,“*=0, (7)
where z=r/R and $= y4/R2.
The dynamic fluid pressure acting on the shell surface is obtained from the Bernoulli
equation

PF = -(pFlps)(r/4s2)(a~lat) at h= 1, (8)
where pF is the density of the fluid.
At the shell-fluid interface, the normal fluid velocity must equal the normal shell
velocity. This implies that
$,* = Z-$(aw/Jt) at a= 1. (9)
Further, for a fluid-filled shell, the following restriction must be imposed:

$., = 0 at k = 0. (LO)

2.3. MODAL EQUATIONS

From previous investigations [6-l l] on modal solutions for non-linear vibrations of


cylindrical shells in a vacuum it is observed that, in order to obtain a consistent solution,
the sum of modal functions for the displacements must satisfy all boundary conditions
and continuity requirements and express the non-linear coupling between the vibration
modes. Thus, the analysis must be initiated by generating the appropriate solutions to
equations (3).
Here, the main results of a previous study [14] based on a perturbation procedure are
to be used. This procedure allowed selecting the modes in the assumed displacement
functions and retaining the most significant modes to the first approximation of non-
linearity, in particular the axisymmetric modes (ring modes) which, as pointed out by
Evensen [13] and Ginsberg [lo], play an important role in the outcome of the results.
By the perturbation approach described in reference [14], it can be shown that a
particular solution for the displacement field can be written as follows:
u = C 1 U,(t) cos (in@) cos (jq[) + 1 C Uk,( t) cos (he) sin (lqt), (114
i=1,3,Sj=1,3,S k=0,2,4 1=2,4

v = C 1 y>(t) sin (id) sin (jq&) + C c Vkl(t) sin (kn@) cos (bt), (lib)
1=1,3.Sj=1.3.S k=2,4.6 1=0,2,4

w = 1 C Wi,(r) cos (in@) sin (jqt)+ C 1 wkl(r) cos we) cos (lqt)y (llc)
i=l,3,S j=1,3,S k=0,2,4 1=0.2.4

where UJ r), Vlj(r), etc., are generalized co-ordinates which represent the dependence on
time, n is the number of circumferential waves associated with the linear free vibration
modes and
q = msrR/ L, (12)
in which m is the number of axial half waves in the linear problem (see the Appendix).
NON-LINEAR FLUID-SHELL INTERACTION 137
Here the approximated radial deflection function w, obtained from equation (1 lc), was
chosen to have the following form, which includes both axisymmetric and asymmetric
modes and attends the boundary conditions (5):
w = w0co? ot [cos 2qe - (cos 4qt)/4 - (3/4)] + w, cos wt sin qe cos ntl, (13)
where w is the radian frequency. This form was found, through a perturbation-based
procedure, to be the simplest and the best approximation among the combinations of
trigonometric functions attempted.
By retaining in u and o all the modes that couple with w0 and w, through quadratic
and cubic products of displacements and their spatial derivatives in equations (3) and
by taking into account the boundary conditions (5), one obtains for u and u the following
approximate solutions:
u = u0 cos* wt[sin 2qt - (sin 4qt)/2] + u, cos wt cos qt cos no
+u, cos* ot cos 2nf3 [sin 2q[ - (sin 4qe)/2] + u3 cos3 wt cos nk9cos 3q.f
+uq cod ot cos no cos 5qt+ u5 cos6 ot [sin 6q[- (3 sin 8q[)/4], (14a)
u = v, cos wt sin q[ sin nf3+ u2 cos* wt sin 2ne (cos 2q[- 1)
+ v3 cos3 wt sin ne sin 3qe + vd cos5 or sin ne sin 5q[. (14b)
It is useful to note that these functions and consequently all force and moment resultants
satisfy the circumferential periodicity condition.
The boundary condition (9) suggests seeking the solution for the velocity potential in
the form
6 = A1j,(e)u sin ot sin q.$ cos ne+ A2fi(c)w sin 2wt cos 2q[
+A3f3(Om sin 2wt cos 4q,f+A,f,(r, [)w sin 2wt. (15)
Substituting equation (15) into equation (7) and taking into account the restriction (lo),
one obtains, for the functions A in equation (15),
J-l(e) = I”(CP), _M#) = I,(2qr), M.2) = I0(4@), Jl(% 5) = a2- 2(5* - &), (16)
where I,, is the nth order modified Bessel function of the first kind [19].
Substituting the expressions for w and 6 into equation (9), one can write the amplitudes
Ai of 4 in terms of w1 and w,, as follows:
A, = -2y8w,{q[I,-,(q) -(n/sL(s)l)-‘, A2 = -~~wo/qI,(2q),
(17)
A3 = y%/fhI,(4q), A4 = 3 ySw,,‘4.
Finally, the expressions for the displacements (13) and (14) and for the velocity potential
(15) are substituted into equations (3) and (8) and a Galerkin procedure is used to obtain
12 coupled algebraic non-linear equations for the modal amplitudes Ui, Vi and Wi. The
expressions aw/aw,,au/au, and au/au, were used as the weighting functions in the Galerkin
procedure. These non-linear equations are unfortunately too long to be presented in the
present report; the interested reader will find them in reference [14].
The complexity of the algebraic equations precludes the possibility of closed-form
solutions. Nevertheless, the Newton-Raphson method can be effectively used to solve
these equations in relatively few interactions. Through this procedure one obtains the
relation between the frequency and the vibration amplitudes which allows one to study
the influence of fluid and shell parameters on the degree and type of non-linearity involved
in the frequency response. The plot of a response-frequency relationship is usually referred
to as the “backbone” curve [23].
138 P. B. CONCALVES AND R. C. BATISTA

3. DISCUSSION OF RESULTS
In the analysis of the results the following frequency ratios are introduced:

A, = GJG”, AF = flFI&F. (18)


Here 0, (= yw,) and flnF (= ywF) are the lowest non-linear frequency parameters (associ-
ated with predominantly radial motion) for, respectively, a shell in a vacuum and a
fluid-filled shell, and 0,” and fioF are the corresponding linear free vibration frequencies
(see the Appendix).
In Figure 2 the frequency ratios AF and A, are plotted as functions of the maximum
non-dimensional radial amplitude w,,,( W,,,,,/ h) f or selected values of the circumferential
wavenumber n. It is shown that the degree of non-linearity displayed by these backbone
curves increases as the number of circumferential waves increases. The curves shown in
Figure 2 differ from those that are usually depicted for the cases where either non-linear
softening or hardening terms alone are involved. Initially, the backbone curves bend
toward decreasing frequencies (A < 1) until a vertical tangent occurs at a finite value of
A, thereafter bending back, reversing slope. The initial curves’ behaviour is due to the
influence of both non-linear elasticity terms containing quadratic powers of the generalized
co-ordinates and inertia terms associated with second-harmonic terms which, together,
are predominant at this initial stage. In the next stage, when the curves bend back to
positive slope and proceed indefinitely to the right, the hardening terms associated with
higher order elasticity terms in the modal equations predominate. This reversal in slope
is stronger for shells in a vacuum.

Figure 2. Frequency of vibration us. maximum radial amplitude for different values of the circumferential
wavenumber, n. -, Fluid-filled shell; - - -, shell in vacuum; - . -, fluid-filled shell (1st mode only).
Rf h = 200, L/mR = 0.6, Y = 0.30, pF/ps = 0.132.

The effect of the fluid region on the shell behaviour can be separated into two different
parts: the inertia effects associated with the classical radial mode (w, cos n6 sin 45 cos wt)
and the inertia effects associated with the secondary radial modes (axisymmetric modes).
As shown in Figure 2 for n = 6 (dash and point curve), if only the inertia effects associated
with the classical radial mode are retained, the non-linearity for a fluid-filled shell is
NON-LINEAR FLUID-SHELL INTERACTION 139
weaker than that predicted for a shell in a vacuum. This “added mass” effect is also
responsible for the significant reduction in the linear free vibration frequencies of a
fluid-filled shell, as shown in Figure 3 (L&CC J&J. On the other hand, if the fluid effects
associated with the shell secondary modes are also taken into account, the non-linearity
for a fluid-filled shell, as shown in Figure 2, is much stronger than that predicted for a
shell in a vacuum. This strong softening effect is due to the increase of the inertia effects
associated with second-harmonic terms in the non-linear equations of motion [14,21].
The results indicated in Figures 4 and 5 show how variations in the geometrical
parameters and in the axial wavenumber affect the degree and type of non-linearity of
the vibrations. The curves in Figure 4 demonstrate that, for given values of R/h and n,
the non-linearity increases as the ratio L/R decreases or as the number of axial waves
increases: i.e., the non-linearity varies inversely with the axial wave length.
The curves in Figure 5 were obtained for a fixed L/h ratio and different values of the
radius, R. For each geometry the wavenumbers (m, n) are those associated with the
minimum natural radial frequency in the spectrum. As shown in Figure 5, the degree of
non-linearity associated with the minimum natural frequency increases as the ratio R/h
increases and/or the ratio L/R decreases: i.e., strong non-linearities occur for the case
of thin and/or short cylinders. It was observed during the calculations that the non-linearity
associated with the minimum natural frequency is usually of the softening type, at least

OL
4 6 8 IO 12 14 16
n

Figure 3. Frequency spectrum.-, Fluid-filled shell; - - -, shell in vacuum. R/h = 200, L/ mR = 04, Y = 0.30,
pF/ps = 0.132.

Figure 4. Influence of the axial wavelength on the frequency-amplitude relationship. -, Fluid-filled shell;
- - -, shell in vacuum. R/h = 100, Y = 0.30, pF/ps = 0.132, n = 10.
140 P. B. GONCALVES AND R. C. BATISTA

Figure 5. Influence of the radius-to-thickness


ratio and length-to-radius ratio on the frequency-amplitude
relationship. L/h = 120, ~=0.30, &/ps =0.132,
m = 1. -, R/h =200,L/R =0*6,n = 10; - - -, R/h = 100,
L/R=1.2,n=6;-..-, R/h = 50,L/R = 2.4,n = 4.V, shell in vacuum; F, fluid-filled shell.

for amplitudes up to the order of the shell thickness. This agrees with the few experimental
results reported in the literature for the case of shells in a vacuum [22,23]. For the case
of fluid-filled shells, to the best of the authors’ knowledge, no experimental results are
available in the literature.
For small vibration amplitudes, the present results for the in vacuum case compare
well with those obtained by Evensen [ 131 using Donnell equations. For larger amplitudes,
the Evensen results show a stronger softening behaviour. This difference is to be expected,
since Donnell equations in terms of the radial displacement and a stress function are not
appropriate for genuinely large deflections [27].
Figure 6 illustrates the effect of a compressive prestress state on the non-linear behaviour
of a shell in a vacuum. In Figure 6 the following ratios are introduced:

% = %,/ %CR, u8 = %d %CR* (19)

Here nxCR and nOCRare the linear critical values for, respectively, a shell under axial
loading and a shell under lateral pressure [24]. It can be observed that by loading the
shell in compression the non-linear effects become much more pronounced, especially
for the axial loading case. A similar behaviour is observed for the case of a fluid-filled shell.

Figure 6. Influence of a prestress state on the non-linear behaviour of the shell. R/h = 200,L/mR = 0.6,
Y = 0.30,
n = 10.(a) Axiai load; (b) lateral pressure.
NON-LINEAR FLUID-SHELL INTERACTION 141

i1.5 )

r
2

j
-
I

(I.51

Figure 7. Response to excitation near the fundamental frequency for a fluid-filled shell. R/h = 200, L/mR =
0.6, Y = 0.3, pF/ps = 0.132, n = 10. -, Stable response; - - -, unstable response; - . -, backbone curve.

For the case of forced vibrations, the applied loading FJt, f?, r) was chosen so that
only one mode is directly excited: i.e.,
ijr = G,,,, cos n6 sin q,$ cos wt. (20)

Figure 7 illustrates the response curves for both free and forced vibrations of a fluid-filled
shell. The unstable solution is denoted by a dashed line. The loci of the vertical tangencies
in Figure 7 together with the backbone curve define the boundaries of an instability region
which, as shown by Stoker [25], corresponds to the first instability region of the Mathieu
equation.
One must bear in mind that in a real system dissipation is always present. The presence
of dissipative forces will diminish the instability region and give rise to a rounded peak
on the skewed resonance curve, the location of which depends upon the relative amount
of damping [25,26].
Although it is not discussed here, one can conclude easily that the degree of non-linearity
increases as the ratio (pF/ps) increases.

4. CONCLUDING REMARKS
The effect of a fluid medium on the degree and type of non-linearity of a thin-walled
cylindrical shell has been investigated.
It has been shown for a broad range of values of wavenumbers (m, n) and geometric
ratios (R/L, h/R) that, for excitation at the fundamental frequency, the non-linearity is
initially of a softening type. This initial frequency behaviour is due to the non-linear
modal coupling, which reduces the membrane stiffness of the shell and gives rise to inertia
effects associated with second-harmonic terms.
Regarding the effect of the fluid, it has been observed that a fluid-filled shell exhibits
a stronger non-linearity than a shell in a vacuum. For both cases, the non-linearity
increases whenever the numbers of axial and circumferential waves increase and the
ratios h/R and L/R decrease.
By employing a consistent shell theory, and by obtaining a solution that satisfies the
imposed boundary and continuity conditions and retains the most important secondary
modes, the present analysis overcomes many of the shortcomings encountered in previous
studies for shells in a vacuum. Hence, the present formulation for the shell problem is
142 P. B. GON(;ALVES AND R. <‘.BATISTA

capable of furnishing a consistent solution for a wide range of geometric parameters and
wavenumbers.

REFERENCES
1. A. W. LEISSA 1973 NASA SP-288. Vibration of shells.
2. C. L. DYM 1973 Journal ofSound and Vibration 29,189-205. Some new results for the vibrations
of circular cylinders.
3. Y. MNEV and A. PERTSEV 1971 Wright-Patterson Air Force Base Report FTD-MAT-24-118-71.
Hydro elasticity of shells.
4. M. C. JUNGER and D. FEIT 1972 Sound, Structures and Their Interaction. Cambridge: MIT Press.
5. S. J. BROWN 1982 American Society of Mechanical Engineers Journal of Pressure Vessel Technology
104,2- 19. A survey of studies into the hydrodynamic response of fluid-coupled circular cylinders.
6. H. N. CHU 1961 Journal of Aerospace Sciences 28, 602-609. Influence of large amplitudes on
the flexural vibrations of a thin circular cylindrical shell.
7. D. A. EVENSEN 1963, American Institute of Aeronautics and Astronautics Journal 1, 2857-2858.
Some observations on the nonlinear vibration of thin cylindrical shells.
8. E. H. DOWELL and C. S. VENTRES 1968 International Journal of Solids and Structures 4,
975-991. Modal equations for the nonlinear flexural vibrations of a cylindrical shell.
9. S. ATLURI 1972 International Journal of Solids andStructures 8,549-569. A perturbation analysis
of non-linear free flexural vibrations of a circular cylindrical shell.
10. J. H. GINSBERG 1973 Transactions of the American Society of Mechanical Engineers, Journal
of Applied Mechanics 40,472-477. Large amplitude forced vibrations of simply supported thin
cylindrical shells.
11. H. RADWAN and J. GENIN 1976 Transactions of the American Society of Mechanical Engineers,
Journal of Applied Mechanics 43, 370-372. Non-linear vibration of thin cylinders.
12. J. L. SANDERS, JR. 1963 Quarterly of Applied Mathematics XXI, 21-36. Nonlinear theories of
thin shells.
13. D. A. EVENSEN 1974 in Thin Shell Structures, 133-156. Nonlinear vibrations of circular
cylindrical shells. Englewood Cliffs, NJ. Prentice-Hall.
14. P. B. GONCALVES 1987 Ph.D. Dissertation, COPPE-Federal Unioersity of Rio de Janeiro.
Non-linear dynamic interaction between fluid and thin shells (in Portuguese).
15. J. H. GINSBERG 1975 Joumalof Soundand Vibration 40,359-379. Multi-dimensional non-linear
acoustic wave propagation, Part II: The non-linear interaction of an acoustic fluid and plate
under harmonic excitation.
16. J. H. GINSBERG 1978 Journal of the Acoustical Society of America 64, 1671-1678. Propagation
of nonlinear acoustic waves induced by a vibrating cylinder. I. The two-dimensional case.
17. L. D. LANDAU and E. M. LIFSHITZ 1959 Fluid Mechanics. Reading, Mass.: Addison-Wesley.
18. H. LAMB 1945 Hydrodynamics. New York: Dover.
19. N. W. MCLACHLAN 1934 Bessel Functions for Engineers. Oxford University Press.
20. P. G. GONCALVES and R. C. BATISTA 1987 Journal of Sound and Vibration 113, 59-70.
Frequency response of cylindrical shells partially submerged or filled with liquid.
21. V. V. BOLOTIN 1965 Dynamic Stability of Elastic Systems. San Francisco: Holden-Day.
22. M. D. OLSON 1965 American Institute of Aeronautics and Astronautics Journal 3, 1775-1777.
Some experimental observations on the nonlinear vibration of cylindrical shells.
23. J. C. CHEN and C. D. BABCOCK 1975 American Institute of Aeronautics and Astronautics
Journal 13, 868-876. Nonlinear vibrations of cylindrical shells.
24. D. 0. BRUSH and B. 0. ALMROTH 1975 BucklingofBars, PlatesandShells. Tokyo: McGraw-Hill.
25. J. J. STOKER 1950 Nonlinear Vibrations. New York: Interscience Publishers.
26. H. A. NAYFEH and T. D. MOOK 1979 Nonlinear Oscillations. New York: John Wiley.
27. W. T. KOITER 1963 Proceedings K. Ned. Akad. Wet. Series B 66, 265-279. The effect of
axisymmetric imperfections on the buckling of cylindrical shells under axial compression.

APPENDIX: LINEAR FREE VIBRATION ANALYSIS


For the linear vibrations of a simply supported cylinder without axial constraint the
middle surface displacements can be written as
(u, u, w) = (u, cos q( cos n0, u, sin qe sin n6, w, sin q[ cos no) cos ot. (Al)
NON-LINEAR FLUID-SHELL INTERACTION 143
Considering equations (7), (8) and (9), one obtains for the dynamic fluid pressure acting
on the shell wall [20]

Introducing equations (Al) and (A2) into equations (3) and ignoring the non-linear terms
associated with the incremental displacements as well as the external loads, one obtains
the frequency determinant
[~5,-L!*6,(1+&&~)[=0, i,j=1,2,3. (A3)
Here,
L,,=q*+an*[l+(A/4)], LIZ= Lzl = qn[(3Aa/4) + Q - 11,

L,, = L3, = -q( v -Am*), L2,=n2(1+A)+q2a[l+(9A/4)]+n,,,

L,,= L,,=n[l+An*+Aq*(l+a)+n,,], L,,=1+A(q2+n2)2+n,,n2+n,,q2.


644)

6, denotes the Kronecker delta and R is the dimensionless frequency parameter which
is related to the radian frequency o by
R = Rw[p,(l- V2)/E]1’2. (A5)
The parameter I\;i in equation (A3) may be considered as the “added mass” effect of
the fluid. For a fluid-filled shell it is given by [20]

The determinant (A3), when set equal to zero, yields a bi-cubic equation in the frequency
parameter which yields three frequencies for each pair of wavenumbers (m, n). The lowest
value of R (0,) is associated with predominantly radial motion. If n;i = 0, one obtains
the classical frequency equation for a shell in a vacuum [2].

You might also like