A Spectral Finite Element For Analysis of Wave Propagation in Uniform Composite Tubes

Download as pdf or txt
Download as pdf or txt
You are on page 1of 35

ARTICLE IN PRESS

JOURNAL OF
SOUND AND
VIBRATION
Journal of Sound and Vibration 268 (2003) 429–463
www.elsevier.com/locate/jsvi

A spectral finite element for analysis of wave propagation in


uniform composite tubes
D. Roy Mahapatra, S. Gopalakrishnan*
Department of Aerospace Engineering, Indian Institute of Science, Bangalore 560012, India
Received 14 February 2002; accepted 20 November 2002

Abstract

A spectral finite element model (SFEM) for analysis of coupled broadband wave propagation in
composite tubular structure is presented. Wave motions in terms of three translational and three rotational
degrees of freedom at tube cross-section are considered based on first order shear flexible cylindrical
bending, torsion and secondary warping. Solutions are obtained in wavenumber space by solving the
coupled wave equation in 3-D. An efficient and fully automated computational strategy is developed to
obtain the wavenumbers of coupled wave modes, spectral element shape function, strain–displacement
matrix and the exact dynamic stiffness matrix. The formulation emphasizes on a compact matrix
methodology to handle large-scale computational model of built-up network of such cylindrical
waveguides. Thickness and frequency limits for application of the element is discussed. Performance of
the element is compared with analytical solution based on membrane shell kinematics. A map of the
distribution of vibrational modes in wavelength and time scales is presented. Effect of fiber angle on natural
frequencies, phase and group dispersions are also discussed. Numerical simulations show the ease with
which dynamic responses can be obtained efficiently. Parametric studies on a clamped–free graphite–epoxy
composite tube under short-impulse load are carried out to obtain the effect of various composite
configurations and tube geometries on the response.
r 2003 Elsevier Science Ltd. All rights reserved.

1. Introduction

Metallic tubular structures are used extensively in piping and skeletal components. There are
increasing usage of fiber-wound and laminated composite cylinders and tubes in automobiles,
aircrafts and spacecrafts. For their high strength, high stiffness and light weight, graphite–epoxy
composite strut tubes have been chosen for International Space Station (ISS) Freedom [1]. Due to

*Corresponding author. Tel.: +91-80-309-2757; fax: +91-80-360-0134.


E-mail address: krishnan@aero.iisc.ernet.in (S. Gopalakrishnan).

0022-460X/03/$ - see front matter r 2003 Elsevier Science Ltd. All rights reserved.
doi:10.1016/S0022-460X(02)01539-0
ARTICLE IN PRESS

430 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

high-load carrying ability under pressurized condition, such tubular beams are useful in inflatable
space structures [2]. Significant research has been reported in literature to address issues related to
vibration and noise transmission [3], impact dynamics, fatigue and damage in such structural
components. In problems related to high-frequency vibration, noise transmission and impact,
many analytical and numerical methods based on wave motion have been reported. The present
work deals with development of a new spectral finite element (FE) for efficient analysis of
broadband wave propagation in uniform composite tubes and connected skeletal structures. Also,
special emphasis is made to model coupled wave propagation in such uniform tubular element due
to variation in the angle of laminated composite ply orientation and fiber winding.
Composite tubular structures can be modelled in two ways: (1) The general method is to use
cylindrical shell kinematics. Various theories (e.g., by Mirsky and Hermann [4], Cooper and
Naghdi [5], Greenspon [6], and later by Reddy and Liu [7], Leissa and Chang [8] and Qatu [9] for
composite shell) in this direction have been reported, which are based on the simpler and
computationally tractable framework and observations of infinite order frequency spectral
characteristics from three-dimensional (3-D) analysis [10]. Xi et al. [11] has used 3-D elasticity
solution and shell radial displacement from finite strip element analysis to develop semi-analytical
model to study of characteristic waves, their phase and group dispersions in laminated composite
cylindrical shell. (2) For closed cylindrical shell, often it is useful to represent an equivalent thin-
walled beam kinematics and this latter approximation based on non-axially symmetric first order
shear kinematics is used in the present work.
For uniform circular cylinders, singly curved shell kinematics is sufficient. For coupled wave
propagation analysis, the displacement field in this case requires in-plane displacements, bi-
directional bending and rotation about the shell normal. For a thick shell, transverse shear
deformation becomes significant and 30% or higher error in deflection and natural frequencies
may occur when the effect of such shear is neglected [12]. Considering first order shear
deformation in laminated anisotropic shells, FE results and exact results for simply supported
boundary conditions have been reported in the work of Reddy [13,14], Chandrashekhara [15] and
Sun et al. [12]. In high-frequency vibration analysis, it is essential to consider the wave band, and
representation of the entire modal group is important compared to accuracy of one or two specific
vibration modes. Therefore, analysis in wavenumber space (k space) is found most suitable.
Langley [16,17] has studied modal density and energy flow in cylindrical shells using semi-
numerical analysis. Smith [18] has discussed various types of wave propagation in closed
cylinders. Fuller and Fahy [19] have discussed propagation of axial waves in fluid-filled cylindrical
shell. Recently, frequency-domain methods have been developed by Xi et al. [20], and Ruzzene
and Baz [21] for analysis of longitudinal and circumferential wave motion in fluid-loaded
cylindrical shells. The strip element method developed in Ref. [20] has also been used for studying
the complex wave-mode coupling due to anisotropy and finite shell thickness in laminated
composite [11] and functionally graded cylinders [22,23]. As pointed out in Refs. [3] and [17], the
wave amplitudes for in-plane and flexural (circumferential) motions have significant scale
difference. However, Bennet and Accorsi [24] have shown that in the presence of ring stiffeners,
wave energy flow has stop bands, which are broadband and can be tuned to required frequency
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
band. It can be seen from the expression of ring frequency [17] given by oR ¼ ð1=RÞ E=rð1  n2 Þ
(for flexural wave motion in cylindrical isotropic shell) that natural frequency increases as the
radius of curvature R decreases. This tells us that for closed circular cylinders with shorter span
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 431

and higher thickness-to-radius ratio, the flexural vibrational modes take transition from low- to
high-frequency zone and fall in comparable scale with in-plane motion. In addition, the rotation
about the shell normal causes overall torsional motion in such tubular structure and interacts with
the in-plane and flexural motion depending upon the geometrical parameter Z ¼ h=R; where 2h is
the shell thickness. This length–scale interaction has significance for coupled wave propagation in
laminated or fiber-wound composite tubular structures due to wide scope of tailorability. Further,
for broadband wave propagation (waves with low as well as high group speeds for axial, flexural
and torsional motion in composite tube), a time scale t ¼ Rc1 g characterizes the dynamics, where
cg ð¼ do=dkÞ is the group speed of a particular type of wave. Kaplunov et al. [25] have discussed
the wave motion in thin-walled elastic bodies considering the above scale effects.
Several studies on free and forced vibration of composite thin-walled bodies have been
reported. Song and Liberescu [26] have modelled composite thin-walled closed-section beams
considering non-classical effects such as primary and secondary warping. Rand [27] has carried
out closed-form analysis of thin-walled beams with arbitrary cross-sections and out-of-plane
warping under static loading. Effect of bending-twist coupling and extension-bending coupling for
different lamination angles have been investigated in the above work. Influence of similar
coupling effects on free vibration response of anisotropic thin-walled closed-section beams have
been reported by Armanios and Badir [28] and Dancila and Armanios [29]. Ferrero et al. [30] have
studied uncoupled torsional motion in thin-walled composite beams with mid-plane symmetry. In
the present paper, we consider coupling between axial, flexural and torstional wave modes in first
order shear deformable composite tube in a general form and present computational strategy to
deal with broadband wave propagation.
Although there are many important applications of high-frequency vibration and elastic wave
propagation in composite thin-walled structures in rotorcrafts, turbomachinery, piping and
skeletal structures, efficient and automated modelling strategies and computational simulations
needs further development. In most of these cases, attempts to solve the problems in closed form
become enormously complex. On the other hand, while using standard FE methods, much care is
needed to ensure appropriate mesh and solution scheme using large number of cylindrical shell
elements or thin-walled beam elements to capture higher order vibrational modes. Wang et al. [31]
have developed theoretical solution for orthotropic thick cylindrical shell under impact load based
on finite Hankel transform and Laplace transform and validated the results using axisymmetric
FE model. Frequency-domain-based spectral analysis of wave motion in thin-walled bodies has
been discussed by Doyle [32]. Also, structural acoustics of a complete cylindrical cavity has been
studied here by solving Helmholtz equation in cylindrical co-ordinate. Spectral FEs for coupled
wave propagation in laminated composite beams and related practical issues have been reported
in earlier works of the authors [33,34].
In the present paper, a new FE is developed for uniform circular composite tube with general
fiber orientation. This FE model uses fast Fourier transform (FFT) to uncouple space and time
from the 3-D coupled wave equations and boundary equations. The set of ordinary differential
equations (ODEs) thus obtained are used in Fourier domain to construct the FE model and
system assembly. Apart from handling the coupled wave modes, this model called spectral finite
element model (SFEM) has the following advantages: (1) It allows one to use a single element to
represent an uniform tube in a connected structure due to exact inertial distribution. (2) Under
any transient loading, frequency as well as time responses can be post-processed easily using
ARTICLE IN PRESS

432 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

forward and inverse FFT, which has been integrated with a general purpose code. (3) For a single
uniform tube, the system size is 12  12 and broadband wave propagation analysis in large
network of such structures can be performed with many order smaller system size and lesser
computation time compared to standard FE models and (4) frequency response function (transfer
function) is the by-product of the approach.

2. Linear wave motion in composite tube

Considering the reference X -axis of the cylinder (Fig. 1) passing through the center of the
annular cross-section, the displacement field can be written in terms of three primary
displacements uo ; vo and wo and three cross-sectional rotations yx ; yy and yz at the center as
follows:
" # " #
x x
uðx; y; z; tÞ ¼ uo ðx; tÞ þ z 1 þ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi yy ðx; tÞ þ y 1  pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi yz ðx; tÞ; ð1Þ
y2 þ z2 y2 þ z2

vðx; y; z; tÞ ¼ vo ðx; tÞ  zyx ðx; tÞ; ð2Þ

wðx; y; z; tÞ ¼ wo ðx; tÞ þ yyx ðx; tÞ: ð3Þ


where u; v and w are the longitudinal, lateral and transverse displacements, respectively at a
material point (x; y; z). yx ; yy and yz are the torsional, transverse bending and lateral bending
rotations. x is the normal distance of a material point measured from the mid-plane reference
contour. We assume that any straight line representing the mean diameter of a circular lay-up
remains straight during deformation. This leads to first order shear flexibility of the annular cross-
section. Such shear flexibility can be considered when appropriate shear correction factor is
introduced and will be discussed later. Note that primary warping for a circular contour is zero
and only secondary warping can occur, which can be derived from the formulations given by Song
and Liberescu [26]. As seen in the expression of axial displacement field (uðx; y; z; tÞ) in Eq. (1), the
non-linear terms appear as the combined effect of bending and radial displacement
(circumferential mode) of the reference contour and is termed as ovaling. Also note that in the

Z,w o
s θz
R
Reference Contour
φ

θ 1 Y,vo
R θy
θx
2
L
X,uo
2h

Fig. 1. Co-ordinate system and d.o.f.’s for the spectral FE for uniform composite tube. Two FE nodes are shown with
solid circles.
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 433

above higher order cylindrical bending model, displacement continuity at the mid-plane contour is
ensured. The bending rotations yy and yz are independent of the curvature and assumed constant
throughout the cross-section as in case of Timoshenko beam model. Further details regarding
similar higher order models for thin-walled closed section beams can be found in Refs. [26,28].
The global bending mode has multiplicity of two due to cross-sectional symmetry about Y - and
Z-axis. The circumferential modes consist of antisymmetric and symmetric thickness stretching.
Resonant wavenumbers for the antisymmetric thickness stretching of a complete circular
cylindrical shell simply supported at the ends is k ¼ 2mp=S; S being the arc length, has been
discussed by Langley [17]. The symmetric thickness stretching is a local higher order effect due to
the Poisson ratio and is of significance only for large Z ¼ h=R and is beyond the scope of the
present thin-walled-beam-type modelling. For uncoupled flexural motion of the cylindrical shell
surface, the associated natural frequency is same as the ring frequency (discussed in the
introduction).
In the following and subsequent derivations, small letters with bold face are used to represent
vectors and capital letters with bold face are used to represent matrices. The constitutive relation
in the element co-ordinate system ðX ; Y ; ZÞ is first expressed as
8 9 8 9
>
< sxx >
= >
< exx >
=
%
txz ¼ Q gxz ; ð4Þ
>
: >
; >
: >
;
txy gxy
where
2 32 32 3
1 0 0 Q% 11 0 Q% 16 1 0 0
% ¼6
Q 40 sin f
76
cos f 54 0 Q% 55
76
0 54 0 sin f
7
cos f 5: ð5Þ
0 cos f sin f Q% 16 0 Q% 66 0 cos f sin f
f is the polar angle of a material point in the cross-sectional plane YZ (shown in Fig. 1).
Expression for the elements of the matrix Q % are obtained from the elasticity matrix C (given in
Ref. [35]) for transversely orthotropic plies in fiber-local co-ordinate system, then by rotating in
the ply-local system (x; s; x) and then imposing plane-stress condition on the thin-walled surface
(radius of curvature R) as sxx ¼ 0; tsx ¼ 0 and txx ¼ 0 as discussed in Ref. [27]. This gives rise to
the matrix Q;% whose elements can be expressed as
1 1
Q% 11 ¼ Q11  Q12 ðQ13 Q23  Q33 Q12 Þ þ Q13 ðQ22 Q13  Q12 Q23 Þ; ð6Þ
D D

1 Q245
Q% 16 ¼ Q16 þ Q26 ðQ12 Q33  Q13 Q23 Þ; Q% 55 ¼ Q55  ; ð7Þ
D Q44

1
Q% 66 ¼ Q66 þ Q26 ðQ26 Q33  Q23 Q36 Þ; D ¼ Q223  Q33 Q22 ; ð8Þ
D
and
Q ¼ GT CG; ð9Þ
ARTICLE IN PRESS

434 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

is the transformation of the elasticity matrix C from the fiber-local co-ordinates to the ply-local
co-ordinates, where y (shown in Fig. 1) represents the fiber orientation in the ply-local co-ordinate
system.
Shear moduli Q% 44 ; Q% 55 and Q% 66 are multiplied by a shear correction factor ðK 0 Þ; which can be
computed from the expression proposed by Cowper [36] as

6ð1 þ nÞð1 þ h02 Þ2 Rh


K0 ¼ ; h0 ¼ ; ð10Þ
ð7 þ 6nÞð1 þ h02 Þ2 þ ð20 þ 12nÞh02 Rþh

where n ¼ n23 ; n13 and n12 corresponding to Q% 44 ; Q% 55 and Q% 66 ; respectively. In the process of
deriving the governing equations of motion and force boundary equations, we come across many
higher order stiffness and mass coefficients, which in compact notation, are expressed,
respectively, as
Z 2p Z Rþh Z 2p Z Rþh
Ajl ¼ % T
Qjl C C r dr df; M ¼ rCT C r dr df; ð11Þ
0 Rh 0 Rh

where
C ¼ f1 y z y% z% y;
% y y;
% z z%;y z%;z g; ð12Þ

y% ¼ yRðy2 þ z2 Þ1=2 ; z% ¼ zf2  Rðy2 þ z2 Þ1=2 g; ð13Þ

y ¼ r cos f; z ¼ r sin f; r ¼ R þ x: ð14Þ

Note that in Eq. (11), Ajl is a 9  9 matrix for each of Q% jl ; and hence we shall use additional two
subscripts after jl while expressing a single stiffness coefficient in the following derivations. The
explicit forms of Ajl and M are given in the appendix. Now, the six wave equations derived using
Hamilton’s principle pertaining to six primary displacement variables can be arranged as follows:

duo : 0 ¼ M11 u. o þ M15 y. y þ M14 y. z  A1111 uo;xx  A1611 vo;xx  A1511 wo ;xx
 ðA1512  A1613 Þyx ;xx A1115 yy ;xx A1114 yz ;xx ðA1519 þ A1618 Þyy ;x
 ðA1517 þ A1616 Þyz ;x ; ð15Þ

dvo : 0 ¼ M11 v.o  M13 y. x  A1611 uo;xx  A6611 vo;xx  A5611 wo;xx
 ðA5612  A6613 Þyx ;xx A1615 yy ;xx A1614 yz ;xx ðA5619 þ A6618 Þyy ;x
 ðA5617 þ A6616 Þyz ;x ; ð16Þ

dwo : 0 ¼ M11 w. o þ M12 y. x  A1511 uo;xx  A5611 vo;xx  A5511 wo;xx


 ðA5512  A5613 Þyx ;xx A1515 yy ;xx A1514 yz ;xx ðA5519 þ A5618 Þyy ;x
 ðA5517 þ A5616 Þyz ;x ; ð17Þ
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 435

dyx : 0 ¼ ðM22 þ M33 Þy. x  M13 v.o þ M12 w. o  ðA1512  A1613 Þuo;xx
 ðA5612  A6613 Þvo;xx  ðA5512  A5613 Þwo;xx
 ðA5522  2A5623 þ A6633 Þyx ;xx
 ðA1525  A1635 Þyy ;xx ðA1524  A1634 Þyz ;xx
 ðA5529 þ A5628  A5639  A6638 Þyy ;x
 ðA5527 þ A5626 A5637  A6636 Þyz ;x ; ð18Þ

dyy : 0 ¼ M55 y. y þ M15 u. o þ M45 y. z  A1115 uo;xx  A1615 vo;xx  A1515 wo;xx
 ðA1525  A1635 Þyx ;xx A1155 yy ;xx A1145 yz ;xx
þ ðA1519 þ A1618 Þuo;x þ ðA5619 þ A6618 Þvo;x þ ðA5519 þ A5618 Þwo;x
þ ðA5529 þ A5628  A5639  A6638 Þyx ;x þðA1648 þ A1549  A1557
 A1656 Þyz ;x þðA5599 þ 2A5689 þ A6688 Þyy
þ ðA5579 A5669 þ A5678 þ A6668 Þyz ; ð19Þ

dyz : 0 ¼ M44 y. z þ M14 u. o þ M45 y. y  A1114 uo;xx  A1614 vo;xx  A1514 wo;xx
 ðA1524  A1634 Þyx ;xx A1145 yy ;xx A1144 yz ;xx
þ ðA1517 þ A1616 Þuo;x þ ðA5617 þ A6616 Þvo;x þ ðA5517 þ A5616 Þwo;x
þ ðA5527 þ A5626  A5637  A6636 Þyx ;x þðA1557 þ A1656  A1549
 A1648 Þyy ;x þðA5579 þ A5678 þ A5669 þ A6668 Þyy
þ ðA5577 þ 2A5667 þ A6666 Þyz ; ð20Þ
and the associated force boundary equations can be obtained as
A1111 uo;x þ A1611 vo;x þ A1511 wo;x þ ðA1512  A1613 Þyx ;x þA1115 yy ;x þA1114 yz ;x
þ ðA1519 þ A1618 Þyy þ ðA1517 þ A1616 Þyz ¼ Nx ; ð21Þ

A1611 uo;x þ A6611 vo;x þ A5611 wo;x þ ðA5612  A6613 Þyx ;x þA1615 yy ;x þA1614 yz ;x
þ ðA5619 þ A6618 Þyy þ ðA5617 þ A6616 Þyz ¼ Vxy ; ð22Þ

A1511 uo;x þ A5611 vo;x þ A5511 wo;x þ ðA5512  A5613 Þyx ;x þA1515 yy ;x þA1514 yz ;x
þ ðA5519 þ A5618 Þyy þ ðA5517 þ A5616 Þyz ¼ Vxz ; ð23Þ

ðA1512  A1613 Þuo;x þ ðA5612  A6613 Þvo;x þ ðA5512  A5613 Þwo;x þ ðA5522
 2A5623 þ A6633 Þyx ;x þðA1525  A1635 Þyy ;x þðA1524  A1634 Þyz ;x þðA5529
þ A5628  A5639  A6638 Þyy þ ðA5527 þ A5626  A5637  A6636 Þyz ¼ Mx ; ð24Þ

A1115 uo;x þ A1615 vo;x þ A1515 wo;x ðA1525  A1635 Þyx ;x þA1155 yy ;x þA1145 yz ;x
þ ðA1559 þ A1658 Þyy þ ðA1557 þ A1656 Þyz ¼ My ; ð25Þ
ARTICLE IN PRESS

436 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

A1114 uo;x þ A1614 vo;x þ A1514 wo;x þ ðA1524  A1634 Þyx ;x þA1145 yy ;x þA1144 yz ;x
þ ðA1549 þ A1648 Þyy þ ðA1547 þ A1646 Þyz ¼ Mz : ð26Þ
Here, /:S;x represents derivative with respect to x: We solve the wave equations (15)–(20) exactly
in frequency domain. However, the numerical procedure involves some iterative approximations
due to unavailability of the closed-form solution to the wavenumbers. This issue is discussed in
detail at the end of this section. The wave type solution thus obtained is finally used to construct
the Fourier domain shape function for semi-infinite and finite tube element. Here, the meaning of
spectral element shape function at a particular frequency (sampling frequency in discrete Fourier
transform (DFT)) is similar to the pseudo-static shape function in time-domain FE at a particular
time step. This concept encompasses the basic framework for the development of SFEM and has
been implemented in the earlier works [33,34]. To proceed further, we consider asymptotic
solution to the primary field vector given by
X
N
uðx; tÞ ¼ u# ðx; on Þ eion t ; ð27Þ
n¼1

where ð#:Þ represents the spectral amplitude. u# ðx; on Þ ¼ fu# o v#o w# o y# x y# y y# z gT represents
pffiffiffiffiffiffiffi the generic
displacement vector along the reference X -axis of the tube element and i ¼ 1: on is the
structural frequency at FFT sampling points and N is the Nyquist point in FFT. In the following
discussions, displacements, strains, stresses, forces, etc. will mean their spectral amplitude unless
explicitly written otherwise. After decoupling, the time harmonic component eion t from the spatial
component in Eq. (27), the spectral amplitudes can be expressed as sum of all the wave modes in
the form eikj x [32], where kj is the wavenumber associated with jth wave mode. Thus, spatial
variation in terms of wavenumbers can be expressed as
u# ðx; on Þ ¼ RL0 u* ¼ T1 ðx; on Þ*u; ð28Þ
where u* represents the wave coefficient vector. L0ð1212Þ is a diagonal matrix with asymptotic
entries in characteristic wavenumbers ðkj ; j ¼ 1; y; 12Þ and can be expressed as
8 ik x
>
> e j if kj is þ ve real;
>
>
< eþikj ðLxÞ if kj is  ve real;
L0jj ¼ ð29Þ
>
> eþikj ðLxÞ
if k is þ ve imaginary;
>
>
j
: ikj x
e if kj is  ve imaginary:
Classification of the exponential entries in Eq. (29) is to keep track of propagating and evanescent
wave components at each FFT sampling frequencies. In Eq. (28), Rð612Þ is defined as the
amplitude ratio matrix and is associated with the wave coefficient vector. This is explained in the
next section. Each of kj in Eq. (29) are obtained as root of the 12th order characteristic equation
(also called dispersion equation):
Det Fðkj Þ ¼ 0; ð30Þ
obtained by substituting Eqs. (27) and (28) in the wave equations. Here, Fðkj Þ is a 6  6 complex
matrix. Although the mathematical representation is exact one, the wavenumbers are computed
numerically as mentioned earlier at each FFT sampling frequency over the broad-frequency
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 437

window. However, for a fully coupled composite configuration arising from thickness-wise
asymmetric fiber winding or lay-up, correlating a particular root to its mode of propagation is not
straight-forward as seen in the complexity in Eqs. (15)–(20). Due to this, significant numerical
error may occur, especially at frequency bands where phase-dispersion cross-overs among
different modes appear. In this work, we adopt a k subspace averaging scheme to speed up
computation with improved accuracy. This involves computation of wavenumbers as in the
partially coupled problems [33] or in other words, involves computation of the wavenumbers
associated with the axial-bending (method reported in Ref. [3]), axial-torsion and torsion-bending
cases separately. Characteristic systems arising from these partially coupled problems are
essentially the submatrices of FðkÞ: We compute the trial roots of Eq. (30) by averaging the
wavenumbers obtained from the k subspaces explained above. Note that the operation is
performed for each FFT sampling frequency on (typically at 512 to 32 768 points depending on
the resolution of the load history). Starting with trial roots to solve Eq. (30) reduces the number of
iterations significantly. For refinement of the roots, IMSLr routine [37] based on Muller’s
method is called in the general purpose spectral FE code, followed by reordering of the
wavenumbers to track individual propagating and evanescent wave modes over a broad frequency
band. In cases where exact solution to the wavenumbers kjn are known, an approximation error in
vibration modes can be obtained as
n
d ¼ j1  eiðkj kj Þh j: ð31Þ
In short-wave regime (wavelength lj ¼ 2p=kj small), where wave amplitudes are in the order of h;
error estimate in dynamics of thin-walled bodies by perturbation of symmetric and antisymmetric
thickness shear modes has been developed by Kaplunov et al. [25]. However, similar numerical
error estimation in the present framework of spectral element development is beyond the scope of
this paper and can be discussed elsewhere.
Wavenumber dispersion curves are now plotted in Fig. 2 for a AS/3501-6 graphite–epoxy
composite tubular cross-section with fiber angle y ¼ 01; h ¼ 0:002 m; Z ¼ h=R ¼ 0:1: The
following material properties are used: elastic moduli: E11 ¼ 144:48 GPa; E22 ¼ 9:632 GPa; E33 ¼
9:412 GPa; G23 ¼ 6:516 GPa; G13 ¼ 7:457 GPa; G12 ¼ 4:128 GPa; the Poissons ratio: n23 ¼ 0:49;
n13 ¼ 0:3; n12 ¼ 0:3 and density: r ¼ 1389:2 kg=m3 : In Fig. 2, only the positive wavenumbers
(forward propagating and evanescent modes) are plotted. Whereas, the negative wavenumbers
with same amplitudes also exist and they represent backward propagating or evanescent modes.
One order-of-scale difference between the wavenumbers associated with propagating axial, shear
and propagating flexural, torsional modes can be seen. The plot also shows that at any material
point, there will be additional propagating shear wave modes due to shear deformations gxz and
gxy above the respective cut-off frequencies. The cut-off frequencies satisfy Det Fð0Þ ¼ 0: Below
these cut-off frequencies, the shear waves are evanescent in nature.

3. Spectral finite element model

To represent all the wave coefficients in terms of only 12 primary unknown coefficients
associated with three translational and three rotational motions, an amplitude ratio matrix Rð612Þ
is introduced in Eq. (28). Here, Rmj u* j represents the entries in actual wave coefficient vector
ARTICLE IN PRESS

438 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

o
u
vo
o
w
200 θ
x
θ
y
θz
150
Re[ kj ]

100

50

0
0
5 0
10 10
20
Im[ kj ] 15 30
20 Frequency (kHz)
40

Fig. 2. Plot of wavenumbers kj for a AS/3501-6 graphite–epoxy composite tubular cross-section with fiber angle y ¼ 01;
h ¼ 0:002 m; Z ¼ h=R ¼ 0:1:

associated with jth mode of propagation (j ¼ 1; y; 12) contributing in mth displacement


component (m ¼ 1; y; 6). The vector Rj ; which forms the jth column of R is obtained by
satisfying the equations
Fðkj ÞRj ¼ 0; ð32Þ

Rmj ¼ 1 8f2ðm  1Þ þ 1 ¼ j or 2ðm  1Þ þ 2 ¼ jg: ð33Þ


After some algebraic manipulations, the above two equations yield the following system of linear
equations to be solved at each on :
2 3
j
F11 ? j
F1ðm1Þ j
F1ðmþ1Þ ? F16j 8 9
6 7>>
>
R1j > >
>
6 ^ & ^ ^ & ^ 7 > >
6 7>>
> ^ > >
>
6 j 7 >
> >
>
6F ? F j
F j
? F j 7< R =
6 ðm1Þ1 ðm1Þðm1Þ ðm1Þðmþ1Þ ðm1Þ6 7 ðm1Þj
6 j 7
6F j j j 7> Rðmþ1Þj >
6 ðmþ1Þ1 ? Fðmþ1Þðm1Þ Fðmþ1Þðmþ1Þ ? Fðmþ1Þ6 7> > >
>
6 7>>
> ^
>
>
>
6 ^ & ^ ^ & ^ 7 >
> >
>
4 5>: R >
;
j j j j 6j
F61 ? F6ðm1Þ F6ðmþ1Þ ? F66
n oT
j j j j
¼  F1m ? Fðm1Þm Fðmþ1Þm ? F6m ; ð34Þ

where Fj ¼ Fðkj Þ: Relative magnitude of different amplitude ratios Rmj (p1) represents whether a
particular wave mode is strong or weak with respect to jth mode. Also, they capture the coupling
among longitudinal, flexural and torsional motions.
We eliminate the 12 unknown wave coefficients in fug * using six displacement boundary and six
force boundary equations. Two nodes, each placed at the center of the tube end cross-sections
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 439

appear sufficient. Therefore, by evaluating Eq. (28) at the element nodes at ð0; 0; 0Þ and ðL; 0; 0Þ;
the element nodal displacement vector can be obtained as
" #
T1 ð0; on Þ
u# ¼
e
u* ¼ T2 u* ; ð35Þ
T1 ðL; on Þ

where the non-singular ð12  12Þ complex matrix T2 represents the local wave characteristics of
displacement field. Combining Eqs. (28) and (35), shape function matrix @e ; which is common in
standard FE is obtained as

u# ðx; on Þ ¼ T1 ðx; on ÞT1


2 u# e ¼ @ðx; on Þe u# e : ð36Þ

The above equation can be used to obtain a strain–displacement matrix Be ; which is a standard
terminology in displacement-based FE approach, and can be derived here as
2 @ @ @ 3
0 0 0 z% y%
6 @x @x @x 7
6 7
8 9 6 7
> e# 6 7
< xx > = 6 @ @ @z% @y% 7
6 7
g# xz ¼ 6 0 0 y 7@ðx; on Þe u# e ¼ Be u# e : ð37Þ
>
: >
; 6 @x @x @z @z 7
g# xy 6 7
6 7
6 7
4 @ @ @z% @y% 5
0 0 z
@x @x @y @y

Note that the exact strain field and stress field can be post-processed efficiently in closed form with
the help of Eq. (37). Next, the relationship between generic displacement vector u# ðx; on Þ and wave
# on Þ from the force
coefficient vector u* in Eq. (28) is used to derive the generic force vector fðx;
boundary conditions (Eqs. (21)–(26)) and is given by

# on Þ ¼ Q0 RL0 u* þ Q1 RL1 u* ;
fðx; ð38Þ

where
2 3
0 0 0 0 ðA1519 þ A1618 Þ ðA1517 þ A1616 Þ
6 7
60 0 0 0 ðA5619 þ A6618 Þ ðA5617 þ A6616 Þ 7
6 7
60 0 0 0 ðA5519 þ A5618 Þ ðA5517 þ A5616 Þ 7
6 ! !7
6 7
Q0 ¼ 6
60 A5529 þ A5628 A5527 þ A5626 7;
7 ð39Þ
6 0 0 0 7
6 A5639  A6638 A5637  A6636 7
6 7
6 7
40 0 0 0 ðA1559 þ A1658 Þ ðA1557 þ A1656 Þ 5
0 0 0 0 ðA1549 þ A1648 Þ ðA1547 þ A1646 Þ
ARTICLE IN PRESS

440 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

2 3
A1111 A1611 A1511 ðA1512  A1613 Þ A1115 A1114
6 7
6 A6611 A5611 ðA5612  A6613 Þ A1615 A1614 7
6 7
6 A5511 ðA5512  A5613 Þ A1515 A1514 7
6 ! ! !7
6 7
Q1 ¼ 6
6 A5522  2A5623 A1525 A1524 7:
7 ð40Þ
6 Sym: 7
6 þA6633 A1635 A1634 7
6 7
6 7
4 A1155 A1145 5
A1144
L1 is a (12  12) diagonal matrix obtained as
@
L1jj ¼ L0jj ; i ¼ 1; y; 12: ð41Þ
@x
By evaluating Eq. (41) at the element nodes placed at ð0; 0; 0Þ and ðL; 0; 0Þ and by substituting u*
using Eq. (35), the element nodal force vector f#e is related to the element nodal displacement
vector u# e as
2 3
ðQ0 RL0 þ Q1 RL1 Þx¼0
6 7 1 # e
f#e ¼ 4 5T2 u ¼ K # e u# e : ð42Þ
ðQ0 RL0 þ Q1 RL1 Þx¼L
# e is a (12  12) complex dynamic stiffness matrix for the spectral FE.
Here, K

3.1. Elements with absorbing boundary: throw-off spectral element

Since the element nodal vectors are obtained explicitly in a systematic manner (Eq. (29)), it
becomes possible to identify the boundary reflection terms in the displacement, strain, stress, etc.
This particular aspect is helpful when one needs to exclude the effect of wave scattering at
structural boundaries or to filter out any noise from nearfield response. In the proposed SFEM,
such an analytical feature of a single-node semi-infinite throw-off spectral element is implemented
by assigning L ¼ N in Eq. (35) and by modifying the amplitude ratio matrix as
2 3
R11 0 R13 0 R15 0 R17 0 R19 0 R1 11 0
6 7
R¼4 ^ 0 ^ 0 ^ 0 ^ 0 ^ 0 ^ 0 5: ð43Þ
R61 0 R63 0 R65 0 R67 0 R69 0 R6 11 0
Use of this throw-off element at the structural boundary amounts to absorption of all the energy
from the structure and hence maximum damping artificially introduced to the response.

3.2. Computational issues

After transforming the element local system (Eq. (42)) into the global structural system and
assembling in the global dynamic stiffness matrix, we need to solve it at each FFT sampling
frequency on : Apart from this major difference compared to time stepping or computation-
intensive modal analysis in standard FE solutions, the proposed SFEM becomes very efficient in
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 441

terms of system size. This is evident from the above derivations that the exact solution to the wave
equation inherited in Eq. (42) allows one to consider a single element over uniform spatial
domain. Hence, broadband wave propagation analysis to obtain frequency responses as well as
time responses in a large structures made of several of similar composite elements becomes highly
automated and computationally much cheaper. However, additional complexities in wavenumber
computation and inversion of T2 using matrix domain decomposition techniques to speed up
computation need further optimization.
Once the global dynamic system is formed with the help of Eq. (42), frequency response
function (FRF) can be computed by inverting the dynamic stiffness matrix directly due to its small
size. We solve the system for unit spectral amplitudes of the load histories
X
N
fðx; tÞ ¼ # on Þ eion t
fðx; ð44Þ
n¼1

at the spectral element nodes. Actual response is computed and post-processed by convolving the
spectral amplitude of the load. Only the dynamic stiffness matrix at each frequency needs to be
stored if repeated analysis is required for various non-stationary load histories. In most general
cases while analyzing a composite network or skeletal structure where the element material
property sets are more than one, wavenumbers can be stored for each material property set and
nodal displacement (as few as the number of global interconnections or structural joints) need to
be stored for any post-processing and other complicated analysis.

3.3. Short- and long-wavelength limits for thin shell and limitations of the proposed model

From the kinematic assumptions (Eqs. (1)–(3)) in the proposed higher order cylindrical
bending model, it is clear that there are restrictions for using this model for wave propagation
in very thin as well as very thick shell. Typically, for thin shell, one would expect the shell
transverse motion as predominant one compared to that due to bending rotation of the shell
cross-section. In thin cylindrical shell, therefore, the axisymmetric radial motion becomes
important. On the other hand, in thick cylindrical shell, one would expect the propagation of
higher order Lamb wave modes (first and higher symmetric stretching modes, and third and
higher antisymmetric modes) apart from propagating longitudinal, flexural and shear wave
modes. For thick cylindrical shell, therefore, the spectral band should be limited below the cut-off
frequencies of the higher order Lamb wave modes that are not included in the kinematics of the
present model. The short- and long-wavelength limits for thin shell (based on Love’s thin-shell
theory), beyond which, significant deviation of the proposed model from the actual behavior may
occur, are discussed below.
Let us consider a cylindrical thin shell segment as shown in Fig. 3, where u; v and w are the
longitudinal, tangential and radial displacements, respectively. By neglecting the effect of bending
moment, transverse shear deformation and rotation of the shell normal, the thin-shell kinematics
[38] can be written as
1 1
exx ¼ u;x ; eff ¼ ðw þ v;f Þ; gxf ¼ v;x þ u;f : ð45Þ
R R
ARTICLE IN PRESS

442 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

dx
x
v u
θ


φ

Fig. 3. Co-ordinate system and d.o.f.’s for laminated composite thin shell (h5R).

The orthotropic constitutive model in ply-local co-ordinate system can be expressed as


8 9 2 38 9
< sxx >
> = Q% 11 Q% 12 Q% 16 > < exx >
=
6 % % % 7
sff ¼ 4 Q12 Q22 Q26 5 eff : ð46Þ
>
: >
; >
: >
;
txf Q% 16 Q% 26 Q% 66 gxf
Substituting Eqs. (45) and (46) in the energy components and applying Hamilton’s principle, the
coupled wave equations for the composite thin shell can be expressed as
 
1 1 2
% % % %
 I0 u. þ A11 u;xx þA12 ðw;x þv;xf Þ þ A26 2 ðw;f þv;ff Þ þ A16 u;xf þv;xx
R R R
 
1 1
þ A% 66 v;xf þ u;ff ¼ 0; ð47Þ
R R
1 1 1
 I0 v. þ A% 12 u;xf þA% 22 2 ðw;f þv;ff Þ þ A% 16 u;xx þA% 26 ðw;x þ2v;xf þu;ff Þ
 R R R
1
þ A% 66 v;xx þ u;xf ¼ 0; ð48Þ
R
1 1 1
I0 w. þ A% 12 u;x þA% 22 2 ðw þ v;f Þ þ A% 26 v;x ¼ 0; ð49Þ
R R R
where
Z þh
ðA% jl ; I0 Þ ¼ ðQ% jl ; rÞ dz: ð50Þ
h

To obtain the characteristic equation, the spectral form of the displacement variables u ¼
fu; v; wgT in k-space can be assumed, which is
X
u¼ u# ðfÞeiðkxon tÞ ; ð51Þ
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 443

where k is the wavenumber in the longitudinal direction. For displacement continuity of the
circumferential motion, u# ðfÞ ¼ u# ð2p þ fÞ: Therefore, for circumferential wave propagation, we
can write
X
u¼ u* ðfÞeiðkxþgfon tÞ ; ð52Þ
where g is the integer wavenumber in tangential direction. The coupling between the longitudinal
and tangential modes for different fiber angles y (Fig. 3) is preserved through the wave coefficient
vector u* : Substituting Eq. (52) in the thin-shell wave Eqs. (47)–(49), the characteristic equation
becomes
Det Gðk; gÞ ¼ 0: ð53Þ
Since our objective here is to draw the restriction of the proposed model as the shell becomes very
thin, h5R; the fundamental axisymmetric modes and pure tangential modes need to be studied in
the limits of short and long wavelengths.
The fundamental axisymmetric modes (longitudinal and radial) and vanishing tangential mode
are recovered from Eq. (53) by substituting Z ¼ 0 and solving for k: In this case, we get a fourth
order characteristic equation in k given by
ak4 þ bk2 þ c ¼ 0; ð54Þ
where
1
a ¼ o2n I0 ðA% 11 A% 66 þ A% 216 Þ þ ðA% 11 A% 22 A% 66  A% 11 A% 226  A% 216 A% 22
R2
þ 2A% 12 A% 16 A% 26  A% 212 A% 66 Þ; ð55Þ
1
b ¼ o4n I02 ðA66 þ A66 Þ  o2n I0 ðA11 A22 þ A22 A66  A212  A226 Þ; ð56Þ
R2
1
c ¼ o6n I03 þ o2n I0 A22 : ð57Þ
R2
In the short-wavelength limit, k-N ) a ¼ 0;
 1=2
ðA11 A22  A212 ÞA66 þ 2A12 A16 A26  A11 A226  A22 A216
) os ¼ o n ¼ ð58Þ
R2 I0 ðA11 A66  A216 Þ
is the frequency at which the wave dispersion has singularity. That is the cylinder experiences
resonance at os in the axisymmetric radial mode. In the long-wavelength limit, k-0 ) c ¼ 0
sffiffiffiffiffiffiffiffi
1 A22
) ol ¼ on ¼ 0; ð59Þ
R I0
is the frequency after which the axisymmetric radial mode again starts propagating and is also
the cut-off frequency for axisymmetric wave propagating in thin shell. For isotropic case,
pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
this becomes ð1=RÞ E=rð1  n2 Þ and is called the ring frequency, which has been discussed in the
pffiffiffiffiffiffiffiffiffi
introductory discussion. Also, for isotropic case, os becomes simply ð1=RÞ E=r and smaller
than the ring frequency. Hence, for isotropic as well as orthotropic materials, it can be said that,
the axisymmetric radial mode first becomes resonant at os and cease to propagate and then again
ARTICLE IN PRESS

444 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

0.5
1
0.4

0.3
kh

0.2

3
0.1

0
0
10

-1
10

η 100
75
-2 50
10 25
0
Frequency (kHz)

Fig. 4. Plot of wavenumbers in axisymmetric modes (1—longitudinal, 3—radial) for different Zð¼ h=RÞ (0.05–0.25) for
y ¼ 01 AS/3501-6 graphite–epoxy composite shell.

Fig. 5. Plot of wavenumbers in axisymmetric modes (1—longitudinal, 3—radial) for different fiber angles y for Z ¼ 0:1
h ¼ 2 mm AS/3501-6 graphite–epoxy composite shell.

starts propagating at little higher frequency, that is at ol : For h ¼ 2 mm; Z ¼ 0:1 as considered in
Fig. 2, these limiting frequencies are os ¼ 20:954 kHz and ol ¼ 21:017 kHz: Fig. 4 shows the
location of the singularity and cut-off in the axisymmetric radial mode (marked 3) in a graphite–
epoxy composite shell with different Zð¼ h=RÞ and y ¼ 0: The non-dispersive longitudinal modes
(marked 1) remain unchanged for all values of Z: Fig. 5 shows the separation between the
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 445

singularity (os on the lower side along the frequency axis) and the cut-off following new
propagation of the axisymmetric radial mode (ol on the higher side along the frequency axis).
Therefore, to exclude the effect of unaccounted radial mode (as in case of the proposed new
tubular element), the best comparable behavior of cylindrical tube (when modelled as beam) can
be obtained for certain range of Z; such that the applied forcing frequency band falls below
os ool : And, as a special case for certain orientation of the fibers, a second frequency band
between os and ol (Fig. 5) can be obtained, within which the axisymmetric radial mode vanishes.
This can be clearly seen from Fig. 6, which shows that almost 35 kHz bandwidth is available with
vanishing axisymmetric radial mode for a tube cross-section having Z ¼ 0:1 and fiber angle y ¼
601: This is the so called stop band for radial mode, which is an important design parameter for
composite shells for controlling vibration and buckling. However, it should be noted that for the
composite tube when designed to behave mainly as a thin-walled beam structure, the beam motion
is less likely to be affected by the radial mode (as the motion of the beam axis remains unaltered),
unless a circumferential normal pressure type loading is applied.
The fundamental cross-sectional warping mode consisting of coupled torsional–radial motion is
recovered by substituting k ¼ 0 in Eq. (53) and solving the fourth order polynomial in g given by

a0 g4 þ b0 g2 þ c0 ¼ 0; ð60Þ

where

1 1
a0 ¼ o2n I0 A22 A66 þ o2n I0 3 A226 ; ð61Þ
R4 R

100

90

80
Limiting frequency (kHz)

70

60

50 (w −w )
l s

40

30

20
w
s
10 w
l

0
0 10 20 30 40 50 60 70 80 90
θ (degree)

Fig. 6. Separation between the short- and long-wave limiting frequencies (ol  os ) for different fiber angles y for
Z ¼ 0:1; h ¼ 2 mm AS/3501-6 graphite–epoxy composite shell.
ARTICLE IN PRESS

446 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

1 1
b0 ¼ o4n I02 ðA22 þ A66 Þ  o2n I0 4 A22 A66 ; ð62Þ
R2 R
1
c0 ¼ o6n I03 þ o4n I02 A22 : ð63Þ
R2
In the short-wavelength limit, g-N ) a0 ¼ 0
A22 A66
) os ¼ on ¼ 0 or R ¼ -N; ð64Þ
A266
which shows that for smaller radius of the cylindrical tube, the wavelength of radial mode
becomes longer and vanishes below the long-wavelength limit ol ; which is also the cut-off
frequency for such propagation. This long-wavelength limit is obtained for g-0 ) c0 -0
sffiffiffiffiffiffiffiffi
1 A22
) ol ¼ on ¼ 0; ; ð65Þ
R I0
which is same as the cut-off frequency of the radial mode in axisymmetric case (Eq. (59)). Fig. 7
shows the decreasing nature of g (hence increasing wavelengths) for the tangential and radial
modes for increasing Z: Also, the cut-off frequency of the radial modes shift towards higher
frequency range for increasing Z (smaller radius for a given shell thickness). Fig. 8 shows similar
shift in the cut-off frequency of radial mode (as in the above case) for increasing fiber angles y for
Z ¼ 0:1: However, the tangential or torsional mode has a symmetry about y ¼ 451 and is non-
dispersive. Such cross-sectional warping is already present in the proposed model (see the
kinematics), where propagation of the radial mode induced by torsional load is not restricted due
to any additional kinematical assumptions.

15

10

2
γ

3
0
100
75 -2
10
50
-1
10
25
Frequency (kHz) η
0 0
10

Fig. 7. Plot of wavenumbers in tangential–radial modes (2—tangential, 3—radial) for different Zð¼ h=RÞ (0.05–0.25)
for y ¼ 01 AS/3501-6 graphite–epoxy composite shell. Only integer values of the wavenumbers g are admissible.
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 447

Fig. 8. Plot of wavenumbers in tangential–radial modes (2—tangential, 3—radial) for different fiber angles y for
Z ¼ 0:1; h ¼ 2 mm AS/3501-6 graphite–epoxy composite shell. Only integer values of the wavenumbers g are admissible.

The above analysis shows that the main limitation of the proposed tubular element while
capturing the behavior of composite thin cylindrical shell is that the axisymmetric radial mode
induced by longitudinal load is absent in the element. Also, the circumferential normal pressure
load on the cylindrical cross-section cannot be modelled. The best comparable thin-shell behavior
can be captured over frequency band below the limiting frequency of short wavelength os and
over the stop bands ðol  os Þ for specified values of Zð¼ h=RÞ and fiber angle y for axisymmetric
modes. On the higher side of Z and smaller L=R; one needs to consider the effect of parabolic
transverse shear stress variation across the shell thickness and vanishing normal stress at the shell
surfaces in the kinematics, which are not included in the present model. Therefore, for accurate
results, application of the present tubular element needs to be restricted for forcing frequency
bandwidth below the lowest of the new cut-off frequencies of any of these un-accounted modes
(i.e., higher order antisymmetric Lamb wave mode due to parabolic transverse shear and
symmetric Lamb wave mode due to vanishing normal stress at the shell surfaces). However, it can
be seen that for both the axisymmetric as well as torsional excitations in very thick cylindrical shell
and rod [10], the lowest longitudinal and torsional modes remain non-dispersive and they are
preserved in the present model. The studies on characteristic wave behavior reported in Ref. [11]
for higher thickness and anisotropy can be adopted to draw the thickness limit for particular
application while using the present SFEM.

3.4. Comparison with analytical solution

Although many theories based on potential functions as well as first order and higher order
shear deformable shell kinematics have been reported in literature as discussed earlier, most of the
wave propagation studies are focused on the analysis of the frequency spectrum and harmonic
analysis in modal space and are difficult to apply for transient dynamic analysis. For validation of
ARTICLE IN PRESS

448 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

the results from the present SFEM, which is mainly suited for broadband and impact-type loading
on composite tubes, we consider the analytical solution for impact-induced response of a semi-
infinite membrane shell [38] for unidirectional (y ¼ 01) composite. The kinematics is given in
Eq. (45). Additional approximations that can be made under longitudinal impact is @=@fð Þ ¼ 0
and v-0: Starting with shear deformable kinematics also, the same approximation remains valid
(except near structural boundaries across the span) under longitudinal impact, and the wave
equations take the form
1
I0 u. þ A11 u;xx þA12 w;x ¼ 0; ð66Þ
R
1 1
I0 w. þ A12u;x þA22 2 w ¼ 0; ð67Þ
R R
’ 0Þ ¼ 0 at t ¼ 0; the boundary conditions u ¼ uðx; tÞ and w ¼
subjected to initial condition uðx;
wðx; tÞ prescribed at a particular x or
1
A11 u;x A12 w ¼ Nx ; ð68Þ
R
where Nx is the applied longitudinal impact. Assuming the solution of the field variables u
transformed into their time uncoupled Fourier coefficients (spectral amplitude) u# as done earlier,
we can write
X % %
u¼ # ion t ; u# ¼ u* 1 eikx þ u* 2 eikx ;
ue ð69Þ
n
X % %
w¼ # ion t ;
we w# ¼ w* 1 eikx þ w* 2 eikx ; ð70Þ
n

where the wavenumber


sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
o2n I0 ð1=R2 ÞA22  o4n I02
k% ¼ ; ð71Þ
o2n I0 A11  ðA11 A22  A212 Þð1=R2 Þ
is obtained by solving the characteristic equation derived from Eqs. (66) and (67). on is the fine
sampling frequency used for forward and inverse FFT and same as the sampling frequency of the
impact loading spectrum N# x ðon Þ (Eq. (44)) used to excite the structure. Considering the free-end
of the semi-infinite complete cylindrical membrane shell under impact load is at x ¼ L and the
other end is at x ¼ N; the displacement spectrum u# finally becomes
% N# x =ð2pRÞ
u# ¼ u* 1 eikx ; u* 1 ¼ % ; ð72Þ
% 11 þ A12 ð1=RÞR21 ÞeikL
ðikA
where
iA12 ð1=RÞk%
R21 ¼ : ð73Þ
o2 I0 þ A22 ð1=R2 Þ
The short-impulse-type loading used for longitudinal impact is shown in Fig. 9. The time duration
is approximately 50 ms with peak amplitude of 100 N: The load spectrum is shown in the inset of
Fig. 9. Fine discretization of the frequency spectrum with Nyquist point N ¼ 16 384 ðDt ¼ 1 msÞ is
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 449

-3
x 10
100 2.5

Frequency Amplitude
2

80 1.5

60
Load (N)

0.5

0
0 50 100 150 200
40 Frequency (kHz)

20

0 50 100 150 200 250 300


Time ( µ sec)

Fig. 9. A short-pulse load history used to excite the composite tubes. Frequency amplitudes at different frequencies are
shown in the inset.

made while inverse FFT of analytically computed velocity spectrum ion u# (Eq. (72)) to obtain the
velocity history at x ¼ L: Graphite–epoxy unidirectional composite (y ¼ 01) with material
properties as considered earlier is used here. The velocity history from analytical result is plotted
in comparison with the velocity history from SFEM (using a single throw-off element as discussed
in Section 3.1) in Fig. 10. It can be seen from the analytical result that the initial impact has
produced non-dispersive longitudinal wave coupled with radial motion after the initial incidence.
Although the tube is semi-infinite along x; the almost stationary axisymmetric radial motion has
non-decaying effect, which has caused window distortion. This numerical problem inherent to the
FFT is evident from the initial non-zero values in the analytical response before the initial
incidence at around 100 ms (Fig. 9). The non-dispersive nature of the longitudinal velocity after
incidence, which is due to coupling with radial motion, could not be captured by the SFEM as the
degrees of freedom (d.o.f.’s) associated with axisymmetric radial motion is absent in the model.
Also, the peak velocity estimate from SFEM is higher than that compared to the analytical result.

4. Numerical simulations

As discussed in the introductory discussion that many solution methods and fundamental
studies have been reported in literature on wave propagation in cylindrical shells. Although few
studies are on composite shells and thin-walled bodies, they are focused on the in-plane and
flexural wave motions separately. However, due to wide tailorability of composite materials
(ratios of elastic moduli in principal co-ordinates, the Poisson ratio, fiber orientations, inclusions,
etc.), a single computational framework is necessary from engineering application point of view.
ARTICLE IN PRESS

450 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

20
analytical
SFEM

15
Longitudinal velocity (mm/sec)

10

-5

-10
0 0.05 0.1 0.15 0.2 0.25
Time (msec)

Fig. 10. Comparison of the mid-plane longitudinal velocity from SFEM and analytical result at the free end of a semi-
infinite graphite–epoxy unidirectional ðy ¼ 01Þ composite membrane shell ðh ¼ 0:002; h=R ¼ 0:1Þ under the short pulse
(Fig. 9) applied in longitudinal direction.

The SFEM presented in the previous section is intended for such purpose. In this section,
numerical simulations are carried out considering a single clamped–free graphite–epoxy tubular
element. Before we analyze the nature of time response of such a structure under short-impulse
load, it is worth looking at the resonant vibration modes among which the energy will be
distributed.

4.1. Modal density distribution in wavelength and time scales

Since the dispersion relation for axial–flexural-torsion coupled wave propagation for the
developed tubular element (uncoupled result is shown in Fig. 2) is not amenable in closed form,
and are fairly complicated due to cut-off frequencies and cross-overs between each other, direct
physical conclusions are difficult to draw. Instead, it appears useful to study the effect of
wavelength and time scales on the dynamics. The importance of modal density distribution in k-
space over the predictions of individual modes in broadband wave propagation problem has been
discussed by Langley [17]. In the present context of coupled wave propagation, we take the notion
of a variability index p and a dynamicity index q as proposed by Kaplunov et al. [25] to
characterize the dynamics. These indices are defined as
lj =R ¼ Zp ; t=ðRc1 q
o Þ¼Z ; ð74Þ
where Z ¼ h=R is a geometric parameter, lj ¼ 2p=kj is the wavelength ofpthe jth wave ffi mode and
ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
t ¼ Rc1
g is the time scale as discussed in the introductory section. co ¼ A 1111 =M 11 is the phase
speed of primary axial wave in the tube as a thin-walled beam. If Z is considered as a length scale
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 451

0.2 0.2

0 0

-0.2 -0.2
Dynamicity index (q)

Dynamicity index (q)


-0.4 -0.4

-0.6 -0.6

-0.8 -0.8

-1 -1

-1.2 -1.2
-4 -3 -2 -1 0 1 2 -4 -3 -2 -1 0 1 2
(a) Variability index (p) (b) Variablity index (p)

Fig. 11. Density distribution of vibrational modes in ðp; qÞ co-ordinates (Eq. (74)) for a clamped–free graphite–epoxy
composite tube with h ¼ 2 mm; Z ¼ 0:1 and fiber angle (a) y ¼ 01; (b) y ¼ 301 over a frequency range of 20 kHz: J;
axial mode; n; transverse bending mode; &; torsional mode: The larger symbols in (b) represent coupled or secondary
resonant modes.

to classify different types of wave, then pX1 represents short-wave region, and subsequently,
high-frequency dynamics can be experienced as q moves from qo1 to q ¼ 1:
To obtain the modal density distribution in (p; q) co-ordinate, first we identify the resonant
frequencies from the location of the poles along frequency axis in the frequency response of
displacements (axial, transverse and torsion) under unit impulse. For uncoupled motions ðy ¼ 01Þ;
all of these poles correspond to the natural frequency of the primary resonant vibrational modes
in individual frequency response. For coupled motions ðya01 in Fig. 1), the coupled or secondary
resonant vibrational modes can be separated from the primary resonant vibrational modes due to
their smaller spectral amplitudes u# : Next, propagating wavelengths and group speeds
corresponding to each of these resonant frequencies are computed to locate them in ðp; qÞ co-
ordinate using Eq. (74). Fig. 11(a) shows the modal density distribution of a clamped–free
graphite–epoxy tube for y ¼ 01; Z ¼ 0:1 and with same material properties as used earlier. It can
be seen from the plot that the resonant axial and torsional vibration modes together represent
always a low-frequency dynamics compared to the increasing nature of the high-frequency
dynamics in resonant bending vibration modes. Higher modal densities of resonant bending and
torsional vibration modes are visible in short-wave region. Fig. 11(b) shows the modal density
distribution for y ¼ 301: Here, it can be seen that the resonant bending and torsional vibration
modes together represent a lower frequency dynamics compared to y ¼ 01 case. On the other
hand, the resonant axial modes moves towards higher frequency dynamics. Interaction in
wavelength scale does not change much. However, the density of primary resonant vibration
modes decreases and a number of secondary modes appear.
ARTICLE IN PRESS

452 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

4.2. Time response under short-impulse load and the effect of fiber orientations

The short-impulse load history used to excite the clamped–free graphite–epoxy composite tube
at its free end is shown in Fig. 9. The time duration is approximately 50 ms with peak amplitude of
100 N: The load spectrum has dominant frequency components up to 40 kHz: Such broadband
also encompasses the propagating shear wave modes (see Fig. 2). We formulated the spectral
element considering undamped system. However, in composite structures, damping is a common
phenomena. Different approaches to include the effect of proportional and non-proportional
damping in SFEM have been proposed in Ref. [34]. In the following simulations, we assume small
amount of damping in the form kj ’kj ð1  iZd Þ; where the damping coefficient (also called loss
factor) Zd ¼ 0:001: Also, improved numerical stability and reduced effect of frequency window
distortion can be achieved by using such artificial damping in an undamped model when accurate
time responses need to be post-processed.
First, we simulate the dynamic response of a clamped–free tube ðL ¼ 1 m; L=R ¼ 50; Z ¼
0:1; y ¼ 01Þ by applying the above short pulse at the free end separately in global X and Z
directions and similar moment of 100 N m about X -axis.
For load applied in X -direction, the deformed outer surface ðx ¼ hÞ geometry is snapped at
t ¼ 0:5 ms in Fig. 12. One limitation of the model is clearly visible from this simulation, that is the
absence of any axisymmetric surface undulations on the axially compressed rings. Such small-
scale effect can be obtained through general shell kinematics as discussed in Section 3.3. For three
different sets of fiber angle, axial displacement and axial velocity histories at the point ð1; 0; R þ hÞ
on the free end are shown, respectively, in Figs. 13 and 14. After the initial incidence of the
impact, repeated reflections from the clamped end of the tube can be seen in both the responses. It
is interesting to note that the y ¼ ½0110 =½90110 configuration, which generates maximum thickness
asymmetry along local x direction, is less responsive compared to y ¼ ½þ45110 =½45110
configuration.

Fixed end

y
x

Fig. 12. Snap of the outer surface geometry at t ¼ 0:5 ms for the graphite–epoxy composite clamped–free tube ðL ¼
1 m; L=R ¼ 50; Z ¼ 0:1 with y ¼ 01 lay-up). A short-pulse load (Fig. 9) is applied uniformly at the free end along X -
direction. Scale-factor for displacement amplification is 1  108 :
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 453

1.5

1
Longitudinal displacement (µm)

0.5

-0.5

-1

-1.5
0 0.25 0.5 0.75 1 1.25 1.5
Time (msec)

Fig. 13. Longitudinal displacement history uðtÞ at a point on the outer top-surface ðy ¼ 0; z ¼ R þ hÞ at the free end
ðx ¼ 1 mÞ of the clamped–free composite tube for different fiber angles: -.-.-, y ¼ ½0120 ; —–, y ¼ ½þ45110 =½45110 ; - - -,
y ¼ ½0110 =½90110 :

90

60
Longitudinal velocity (mm/sec)

30

-30

-60-

-90
0 0.25 0.5 0.75 1 1.25 1.5
Time (msec)

’ at a point on the outer top-surface ðy ¼ 0; z ¼ R þ hÞ at the free end ðx ¼


Fig. 14. Longitudinal velocity history uðtÞ
1 mÞ of the clamped–free composite tube for different fiber angles: -.-.-, y ¼ ½0120 ; —–, y ¼ ½þ45110 =½45110 ; - - -,
y ¼ ½0110 =½90110 :
ARTICLE IN PRESS

454 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

For load applied in Z-direction, the deformed outer surface ðx ¼ hÞ geometry is snapped at
t ¼ 0:5 ms in Fig. 15. The effect of ovaling near the mid-length of the tube can be observed. For
three different sets of fiber angle, the transverse displacement and the transverse velocity histories
at the point ð1; 0; R þ hÞ on the free end are shown, respectively, in Figs. 16 and 17. Unlike axial

Fixed end

Fig. 15. Snap of outer surface geometry at t ¼ 0:5 ms for the graphite–epoxy composite clamped–free tube ðL ¼
1 m; L=R ¼ 50; Z ¼ 0:1 with y ¼ 01 lay-up). A short-pulse load (Fig. 9) is applied uniformly at the free end along the Z-
direction. Scale-factor for displacement amplification is 1  107 :

1.2

1
Transverse displacement (µm)

0.8

0.6

0.4

0.2

-0.2
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
Time (msec)

Fig. 16. Transverse displacement history wðtÞ at a point on the outer top-surface ðy ¼ 0; z ¼ R þ hÞ at the free end
ðx ¼ 1 mÞ of the clamped–free composite tube for different fiber angles: -.-.-, y ¼ ½0120 ; —–, y ¼ ½þ45110 =½45110 ; - - -
y ¼ ½0110 =½90110 :
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 455

40

30
Transverse velocity (mm/sec)

20

10

-10

-20

-30
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
Time (msec)

’ at a point on the outer top-surface ðy ¼ 0; z ¼ R þ hÞ at the free end ðx ¼ 1 mÞ


Fig. 17. Transverse velocity history wðtÞ
of the clamped–free composite tube for different fiber angles: -.-.-, y ¼ ½0120 ; —–, y ¼ ½þ45110 =½45110 ; - - -, y ¼
½0110 =½90110 :

responses shown earlier, here, the responses are dispersive. After the initial incidence of the
impact, one reflection from the clamped end of the tube can be seen in both the responses which
arrive at the measurement point at about t ¼ 1:4 ms: As compared to the axial response, here the
y ¼ ½0120 configuration generates maximum transverse response. There are small distortions in
the predicted responses before the incidence of impact (before t ¼ 50 ms), which is due to
frequency window wrap-around during inverse FFT and is inherent to any analysis in
transformed finite domain.
For torsional load applied about X -axis, the deformed outer surface ðx ¼ hÞ geometry is
snapped at t ¼ 0:5 ms in Fig. 18. It can be seen that the rings through which the torsional waves
have propagated towards the clamped end have bulged almost axisymmetrically which can be
attributed to the combined effect of antisymmetric thickness stretching and rotational inertia of
the tube cross-section. For three different sets of fiber angle, lateral displacement and lateral
velocity histories at the point ð1; 0; R þ hÞ on the free end are shown, respectively, in Figs. 19 and
20. Although the torsional waves are non-dispersive for y ¼ 01 (see the dispersion curve in Fig. 2),
the lateral flexural waves are dispersive and finally the combined effect (Eq. (2)) in the lateral
motion on outer top surface becomes attenuating in nature. Due to this reason, incident peaks
and reflected peaks can be visible only in velocity history (Fig. 20) that has dimensional similarity
with phase velocity of the dispersive waves. The displacement and velocity histories corresponding
to y ¼ ½þ45110 =½45110 configuration are, respectively, 2 and 3 order smaller compared to those
from other two configurations. As in the plots of transverse responses against transverse loading,
ARTICLE IN PRESS

456 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

y
x

Fig. 18. Snap of outer surface geometry at t ¼ 0:5 ms for the graphite–epoxy composite clamped–free tube ðL ¼
1 m; L=R ¼ 50; Z ¼ 0:1 with y ¼ 01 lay-up). A short-pulsed torsional loading of peak amplitude 100 N m (similar to
Fig. 9) is applied uniformly at the free end about X -axis. Scale-factor for displacement amplification is 1  106 :

100

50

0
Lateral displacement (µm)

-50

-100

-150

-200

-250

-300

-350

-400
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
Time (msec)

Fig. 19. Lateral displacement history vðtÞ at a point on the top surface ðy ¼ 0; z ¼ R þ hÞ at the free end ðx ¼ 1 mÞ of
the clamped–free composite tube for different fiber angles: -.-.-, y ¼ ½0120 ; —–, y ¼ ½þ45110 =½45110 ; - - -, y ¼
½0110 =½90110 :

here also small errors due to window distortion in the initial responses (before t ¼ 50 ms) have
occurred. However, these errors are within acceptable range and does not alter the nature of
response. Also, these errors can be eliminated by expanding the length of the frequency window
with appropriate resolution.
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 457

4000

3000

Lateral velocity (mm/sec) 2000

1000

-1000

-2000

-3000

-4000

-5000

-6000
0 0.25 0.5 0.75 1 1.25 1.5 1.75 2
Time (msec)

’ at a point on the outer top surface ðy ¼ 0; z ¼ R þ hÞ at the free end ðx ¼ 1 mÞ of


Fig. 20. Lateral velocity history vðtÞ
the clamped–free composite tube for different fiber angles: -.-.-, y ¼ ½0120 ; —–, y ¼ ½þ45110 =½45110 ; - - -, y ¼
½0110 =½90110 :

4.3. Parametric study of impact-induced response

Here we present two parametric studies on a clamped–free graphite–epoxy composite tube, as


considered earlier, with varying geometry and varying fiber angle. The objective is to obtained the
nature of variation of the maximum response due to incident impact at the free-end node for
different L=R and Z: For this purpose, the measure for a particular fiber angle y is chosen as
maximum displacement normalized with respect to maximum displacement for y ¼ 0: Here all the
layers are considered to have same y: In the first type of parametric variation, Z ¼ 0:1 is fixed and
L=R is varied from 20 to 100. For each of these geometries, normalized maximum displacements
are computed for y ¼ 0–901 in steps of 101: In the second type of parametric variation, L=R ¼ 50
is fixed and Z is varied from 0.25 to 0.05. For each of these geometries, normalized maximum
displacements are computed for y ¼ 01–901 in steps of 101: The plots are shown in Figs. 21–26.
From plots of log10 ½uomax =uomax ðy ¼ 01Þ in Figs. 21 and 22, it can be seen that maximum axial
displacement occurs for y ¼ 301 irrespective of the geometry. At this fiber angle, amplitude level
first increases for increasing L=R up to 60 and again decreases towards L=R ¼ 100: From plots of
log10 ½womax =womax ðy ¼ 01Þ in Figs. 23 and 24, it can be observed that maximum displacements
occur at y ¼ 101 irrespective of geometry. Although, relative change in amplitude level when
compared to y ¼ 01 for individual cases vary arbitrarily at different y; one order difference can be
found when compared to those for axial motion in Figs. 21–26. From plots of
log10 ½yox max =yox max ðy ¼ 01Þ in Figs. 25 and 26, it can be seen that the behavior for varying L=R
as well as varying Z is similar to those in axial motion and relative change in the amplitude level
ARTICLE IN PRESS

458 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

2.5
log10[uomax/uomax(θ = 0o)]

1.5

0.5

0
0 10 20 30 40 50 60 70 80 90
Fiber angle θ (degree)

Fig. 21. Plot of log10 ½uomax =uomax ðy ¼ 01Þ at the free-end cross-section of the clamped–free composite tube for various
L=R and various fiber angles y; Z ¼ 0:1: —–, L=R ¼ 20; - - -, L=R ¼ 40; -.-.-, L=R ¼ 60; -.-.-, L=R ¼ 80; - - -, L=R ¼ 100:

2.5
log10[uomax/uomax(θ = 0o)]

1.5

0.5

0
0 10 20 30 40 50 60 70 80 90
Fiber angle θ (degree)

Fig. 22. Plot of log10 ½uomax =uomax ðy ¼ 01Þ at the free-end cross-section of the clamped–free composite tube for various
R=t and various fiber angle y; L=R ¼ 50: —–, Z ¼ 0:25; - - -, Z ¼ 0:125; -.-.-, Z ¼ 0:0833; -.-.-, Z ¼ 0:0625; - - -, Z ¼ 0:05:
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 459

0.6

0.4

0.2
0 )]
o
/w (θ =

0
max max
o

-0.2
o
log [w
10

-0.4

-0.6

-0.8
0 10 20 30 40 50 60 70 80 90
Fiber angle θ (degree)

Fig. 23. Plot of log10 ½womax =womax ðy ¼ 01Þ at the free end cross-section of the clamped–free composite tube for various
L=R and various fiber angle y; Z ¼ 0:1: —–, L=R ¼ 20; - - -, L=R ¼ 40; -.-.-, L=R ¼ 60; -.-.-, L=R ¼ 80; - - -, L=R ¼ 100:

0.6

0.4
log [wmax/wmax(θ =0 )]

0.2
o

0
o
o

-0.2
10

-0.4

-0.6

-0.8
0 10 20 30 40 50 60 70 80 90
Fiber angle θ (degree)

Fig. 24. Plot of log10 ½womax =womax ðy ¼ 01Þ at the free-end cross-section of the clamped–free composite tube for various
R=t and various fiber angle y; L=R ¼ 50: —–, Z ¼ 0:25; - - -, Z ¼ 0:125; -.-.-, Z ¼ 0:0833; -.-.-, Z ¼ 0:0625; - - -, Z ¼ 0:05:

with respect to y ¼ 01 is also similar. Maximum rotation of the tube cross-section occurs near
y ¼ 301 irrespective of the geometry. Also, it can be observed in Fig. 26 that the effect of variation
of Z below y ¼ 101 and above y ¼ 401 is insignificant.
ARTICLE IN PRESS

460 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

2.5

2
(θ = 0 )]
o

1.5
10 x max x max

1

log [θ

0.5

-0.5
0 10 20 30 40 50 60 70 80 90
Fiber angle θ (degree)

Fig. 25. Plot of log10 ½yox max =yox max ðy ¼ 01Þ at the free-end cross-section of the clamped–free composite tube for various
L=R and various fiber angle y; Z ¼ 0:1: —–, L=R ¼ 20; - - -, L=R ¼ 40; -.-.-, L=R ¼ 60; -.-.-, L=R ¼ 80; - - -, L=R ¼ 100:

2.5

2
(θ = 0 )]

1.5
o
10 x max x max

1

0.5
log [θ

-0.5
0 10 20 30 40 50 60 70 80 90
Fiber angle θ (degree)

Fig. 26. Plot of log10 ½yox max =yox max ðy ¼ 01Þ at the free-end cross-section of the clamped–free composite tube for various
R=t and various fiber angle y; L=R ¼ 50: —–, Z ¼ 0:25; - - -, Z ¼ 0:125; -.-.-, Z ¼ 0:0833; -.-.-, Z ¼ 0:0625; - - -, Z ¼ 0:05:
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 461

5. Conclusions

A spectral finite element model (SFEM) for analysis of coupled broadband wave propagation
in uniform composite tube is presented. Related computational complexities, advantages and
limitations of the model for impact induced wave propagation are also discussed. The systematic
formulation in wavenumber space integrated with standard FE approach presented here shows
the possibility of automated wave propagation analysis in more complex composite cylindrical
shells and connected structures. Apart from this main scope of discussions in the paper, numerical
simulations using the new element are performed. In this portion of the work, a map of the
distribution of vibrational modes and their effect on overall dynamics in wavelength and time
scales have been discussed. The approach developed here is for general purpose computational
simulation and does not depend on availability of closed-form expressions to obtain mode count.
Under impact type loading, various time responses and post-processing capabilities are illustrated.
Much higher computational efficiencies compared to standard FE model can be achieved using
this SFEM while solving vibration and wave propagation problems in large connected tubular
structures. Finally, attempt is made to capture the effects of tube geometry and fiber angle on the
impact-induced response of clamped–free graphite–epoxy composite tube. The simulation shows
that the cantilever structure under tip impact load has maximum response for fiber angle y ¼ 301
for axial and torsional motions and fiber angle y ¼ 101 for transverse motion. More detailed
parametric studies on similar aspects along with general higher-order shell kinematics can provide
important inputs for improved design of such slender composite tubes operating in transient
dynamic environment.

Appendix

Z 2p Z Rþh
ðAjl ; MÞ ¼ % jl ; rÞ
ðQ
0 Rh
2 3
1 y z y% z% %y
y; %z
y; z%;y z%;z
6 7
6 y2 yz yy% yz% %y
yy; %z
yy; yz%;y yz%;z 7
6 7
6 z2 zy% zz% zy;%y zy;%z zz%;y zz%;z 7
6 7
6 7
6 y% 2 y% z% y% y;
%y y% y;
%z y% z%;y y% z%;z 7
6 7
6 z%2 %y
z%y; %z
z%y; z%z%;y z%z%;z 7
6 7r dr df;
6 7
6
6
sym % 2y
y; % y y;
y; %z % y z%;y
y; % y z%;z 7
y;
7
6 7
6 % 2z
y; % z z%;y
y; % z z%;z 7
y;
6 7
6 z%;2y z%;y z%;z 7
4 5
z%;2z
where
% 11 ¼ Q% 11 ;
Q % 15 ¼ Q% 16 cos f;
Q % 16 ¼ Q% 16 sin f;
Q
ARTICLE IN PRESS

462 D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463

% 55 ¼ Q% 55 sin2 f þ Q% 66 cos2 f;
Q % 56 ¼ ðQ% 55  Q% 66 Þ sin f cos f;
Q

% 66 ¼ Q% 66 sin2 f þ Q% 55 cos2 f:
Q

References

[1] W.H. Prosser, M.R. Gorman, J. Dorighi, Extensional and flexural waves in a thin-walled graphite/epoxy tube,
Journal of Composite Materials 26 (14) (1992) 418–427.
[2] W.B. Fitchter, A theory for inflated thin-wall cylindrical beams, NASA Technical Note TN D-3466, 1966.
[3] S. Gopalakrishnan, D. Roy Mahapatra, Active control of structure-borne noise in helicopter cabin transmitted
through gearbox support strut, in: M.L. Munjal (Ed.), Proceedings of IUTAM Symposium on Designing for
Quietness, Solid Mechanics and its Applications, Vol. 102, Kluwer Academic Publishers, Dordrecht, 2002.
[4] I. Mirsky, G. Herrmann, Nonaxially symmetric motions of cylindrical shells, Journal of the Acoustical Society of
America 29 (1957) 1116–1123.
[5] R.M. Cooper, P.M. Naghdi, Propagation of nonaxially symmetric waves in elastic cylindrical shells, Journal of the
Acoustical Society of America 29 (1957) 1365–1372.
[6] J.E. Greenspon, Vibration of a thick-walled cylindrical shell—comparison of the exact theory with the
approximate theories, Journal of the Acoustical Society of America 32 (1960) 571–578.
[7] J.N. Reddy, C.F. Liu, A higher-order shear deformation theory of laminated elastic shells, International Journal of
Engineering Science 23 (1985) 440–447.
[8] A.W. Leissa, J. Chang, Elastic deformation of thick, laminated composite shallow shells, Composite Structures 35
(1996) 153–170.
[9] M.S. Qatu, Accurate equations for laminated composite deep thick shells, International Journal of Solids and
Structures 36 (1999) 1917–2941.
[10] D.C. Gazis, Three dimensional investigation of the propagation of waves in hollow circular cylinders—I.
Analytical foundation II. Numerical results, Journal of the Acoustical Society of America 31 (1959) 568–578.
[11] Z.C. Xi, G.R. Liu, K.Y. Lam, H.M. Shang, Dispersion and characteristic surfaces of waves in laminated
composite circular cylindrical shells, Journal of the Acoustical Society of America 108 (5) (2000) 2179–2186.
[12] G. Sun, P.N. Bennett, F.W. Williams, An investigation on fundamental frequencies of laminated circular cylinders
given by shear deformable finite element, Journal of Sound and Vibration 205 (3) (1997) 265–273.
[13] J.N. Reddy, Bending of laminated anisotropic shells by a shear deformable finite element, Composite Science and
Technology 17 (1982) 9–24.
[14] J.N. Reddy, Exact solutions pf moderately thick shells, Journal of Engineering Mechanics Division, American
Society of Civil Engineers 110 (1984) 794–809.
[15] K. Chandrashekhara, Free vibrations of anisotropic laminated doubly curved shells, Computers and Structures 33
(1989) 435–440.
[16] R.S. Langley, Wave motion and energy flow in cylindrical shells, Journal of Sound and Vibration 169 (1) (1994)
29–42.
[17] R.S. Langley, The modal density and mode count of thin cylindrical and curved panels, Journal of Sound and
Vibration 169 (1) (1994) 43–53.
[18] P.W. Smith, Phase velocity and displacement characteristics of free waves in thin cylindrical shells, Journal of the
Acoustical Society of America 27 (1955) 1065–1072.
[19] C.R. Fuller, F.J. Fahy, Characteristics of wave propagation and energy distribution in cylindrical elastic shells
filled with fluid, Journal of Sound and Vibration 81 (1981) 501–518.
[20] Z.C. Xi, G.R. Liu, K.Y. Lam, S.M. Shang, A strip element method for analyzing wave scattering by a crack in a
fluid-filled composite cylindrical shell, Composite Science and Technology 60 (2000) 1985–1996.
[21] M. Ruzzene, A. Baz, Active control of wave propagation in periodic fluid-loaded shells, Smart Materials and
Structures 10 (2001) 893–906.
ARTICLE IN PRESS

D. Roy Mahapatra, S. Gopalakrishnan / Journal of Sound and Vibration 268 (2003) 429–463 463

[22] X. Han, G.R. Liu, Z.C. Xi, K.Y. Lam, Transient waves in a functionally graded cylinder, International Journal of
Solids and Structures 38 (2001) 3021–3037.
[23] X. Han, G.R. Liu, Z.C. Xi, K.Y. Lam, Characteristics of waves in a functionally graded cylinder, International
Journal of Numerical Methods in Engineering 53 (2002) 653–676.
[24] M.S. Bennett, M.L. Accorsi, Free wave propagation in periodically ring stiffened cylindrical shells, Journal of
Sound and Vibration 171 (1) (1994) 49–66.
[25] J.D. Kaplunov, L.Yu. Kossovich, E.V. Nolde, Dynamics of Thin Walled Elastic Bodies, Academic Press, London,
1998.
[26] O. Song, L. Liberescu, Structural modeling and free vibration analysis of rotating composite thin-walled beams,
Journal of the American Helicopter Society, Oct (1997) 358–369.
[27] O. Rand, Fundamental closed-form solutions for solid and thin-walled composite beams including a complete out-
of-plane warping model, International Journal of Solids and Structures 35 (21) (1998) 2775–2793.
[28] E.A. Armanios, A.M. Badir, Free vibration analysis of anisotropic thin-walled closed-section beams, American
Institute of Aeronautics and Astronautics Journal 33 (10) (1995) 1905–1910.
[29] D.S. Dancila, E.A Armanios, The influence of coupling on the free vibration of anisotropic thin-walled closed-
section beams, International Journal of Solids and Structures 35 (23) (1998) 3105–3119.
[30] J.F. Ferrero, J.J. Barrau, J.M. Segura, B. Castanie, M. Sudre, Torsion of thin-walled composite beams with
midplane symmetry, Composite Structures 54 (1) (2001) 111–120.
[31] X. Wang, K. Zhang, W. Zhang, J.B. Chen, Theoretical solution and finite element solution for an orthotropic thick
cylindrical shell under impact load, Journal of Sound and Vibration 236 (1) (2000) 129–140.
[32] J.F. Doyle, Wave Propagation in Structures, Springer, NY, 1997.
[33] D. Roy Mahapatra, S. Gopalakrishnan, T.S. Sankar, Spectral-element-based solutions for wave propagation
analysis of multiply connected laminated composite beams, Journal of Sound and Vibration 237 (5) (2000)
819–836.
[34] D. Roy Mahapatra, S. Gopalakrishnan, A spectral finite element model for analysis of axial–flexural-shear
coupled wave propagation in laminated composite beams, Composite Structures 59 (1) (2003) 67–88.
[35] J.N. Reddy, Mechanics of Laminated Composite Plates, CRC Press, Boca Raton, FL, 1997.
[36] G.R. Cowper, The shear coefficient in Timoshenko’s beam theory, Journal of Applied Mechanics 33 (1996)
335–340.
[37] IMSL Mathematics and Statistics Libraries, Visual Numerics, Inc., San Ramon, CA, 2001.
[38] K.F. Graff, Wave Motion in Elastic Solids, Dover, NY, 1975.

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy