Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Journal of Reinforced Plastics and

Composites
http://jrp.sagepub.com

Analysis of Underwater Free Vibrations of a Composite Propeller Blade


H.J. Lin and J.F. Tsai
Journal of Reinforced Plastics and Composites 2008; 27; 447 originally published online Jan 31,
2008;
DOI: 10.1177/0731684407082539

The online version of this article can be found at:


http://jrp.sagepub.com/cgi/content/abstract/27/5/447

Published by:

http://www.sagepublications.com

Additional services and information for Journal of Reinforced Plastics and Composites can be found at:

Email Alerts: http://jrp.sagepub.com/cgi/alerts

Subscriptions: http://jrp.sagepub.com/subscriptions

Reprints: http://www.sagepub.com/journalsReprints.nav

Permissions: http://www.sagepub.co.uk/journalsPermissions.nav

Citations http://jrp.sagepub.com/cgi/content/refs/27/5/447

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
Analysis of Underwater Free Vibrations of
a Composite Propeller Blade

H. J. LIN*
Department of Electrical Engineering, National Penghu University, Taiwan

J. F. TSAI
Department of Engineering Science and Ocean Engineering
National Taiwan University, Taiwan

ABSTRACT: The free vibration characteristics of a composite marine propeller blade were studied.
MAB and composite materials were used. Composites with symmetric, balanced and unbalanced
stacking sequences were analyzed. Rotational effects and added mass were considered using the
finite element method. The natural frequency of the blade in water was much lower than in air.
The mode shapes of the blade are almost the same in air as in water. The anisotropy of the
composites shift the contours of the mode shape. Generally, greater anisotropy corresponds to
a lower natural frequency. The rotational effects can be ignored because the marine propeller
rotates slowly.

KEY WORDS: underwater, free vibration, composite, propeller blade.

INTRODUCTION

OMPOSITE MATERIALS ARE used in numerous structural applications. They are used
C not only in industry, such as in the mobile, aerospace and shipbuilding industries,
but also in daily life, since they are strong, rigid, lightweight and inexpensive. Composite
materials may now be effectively applied to produce propeller blades. The applica-
tion of composite materials technology to marine architecture has increased with
particular benefits of weight and special characteristics of the materials. The application
of composites may achieve weight, noise and pressure fluctuation reduction and increase
the fuel efficiency. The performance of a composite marine propeller blade involves its
structural, fluid, acoustic, vibrational, material, and other characteristics. Now, many
small or middle size composite marine propellers were commercialized or tested. Thus, the
structural analysis including vibration of the composite marine propeller may be necessary
in the future. The composite marine propeller may be the alternative to metal.
Performing a structural analysis of a propeller blade is difficult because it has complex
geometry and loading. Classical curved beam, plate and shell theories have applied to
analyze the structural analysis of a propeller during the early age [1–3]. The propeller blade
is considered to be a cantilever rigidly attached to a boss. These approaches have been

*Author to whom correspondence should be addressed. E-mail: hjlin@ntu.edu.tw

Journal of REINFORCED PLASTICS AND COMPOSITES, Vol. 27, No. 5/2008 447
0731-6844/08/05 0447–12 $10.00/0 DOI: 10.1177/0731684407082539
ß SAGE Publications 2008
Los Angeles, London, New Delhi and Singapore
Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
448 H.J. LIN AND J.F. TSAI

shown to be successful, but particular assumptions restrict their application. The finite
element method is so popular that various varieties of elements have been used [4–8]. Plate
elements, thin and thick-shell elements and solid elements have been extensively adopted.
Surface loading has been calculated based on assumed loadings or on analytical results
concerning for example, the lifting line, the lifting surface and panel analysis. Coupled
fluid-structure analysis [9] has also been performed to evaluate the strength of composite
propeller blades. An iterative procedure was applied herein to calculate new blade
geometry and modify the surface loading. This process is repeated until stability is once
again achieved. Numerous researchers have performed the free vibration analysis of a
blade in air or in vacuum. However, such a blade is really operated in water, so the
added mass effects due to the water must be considered in the free vibration analysis.
The added mass effect comes from the transmission of pressure to the hull due to the
inertia of the water. Various methods of calculating the added mass are available. They
include, for example, the finite element method, the strip method and ideal flow theory.
This study considers the potential flow of an ideal fluid. A so-called panel method
is adopted, in which the source elements are distributed on the surface of the body
to simulate the flow field. The added mass can be calculated from the surface pressure
induced by the fluid.

GEOMETRY AND STACKING SEQUENCE OF BLADE

The basic blade geometric data of a typical propeller blade are the number of blades (N),
the diameter of the propeller (D or 2R), the radius of the section (r), the length of the chord
(C), the pitch (P), the pitch angle (’), the skew angle (), the rake (Z), the camber (f), the
maximum thickness of the section (t) and the offset of the blade surface. Figure 1 shows
the notation used herein: the Z-axis is the direction of advance; S is a non-dimensional

C/2
C/2
qm Zm
t(s) r 1
s f

0
f(s)

Figure 1. Coordinate system and definitions of variables.

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
Analysis of Underwater Free Vibrations of a Composite Propeller Blade 449

curvilinear coordinate along the nose-tail helix, which are zero at the leading edge and one
at the trailing edge. The coordinates of a point on the face or back surface of the blade
can be written as:

 ¼ m þ C  ðS  1=2Þ  cosðÞ=r þ ð f  t=2Þ  sinðÞ=r ð1Þ

X ¼ r  cosðÞ ð2Þ

Y ¼ r  sinðÞ ð3Þ

Z ¼ Zm þ C  ðS  1=2Þ  sinðÞ  ð f  t=2Þ  cosðÞ: ð4Þ

The propeller selected in the example is the MAU3-60 propeller designed for a single-
screw fishing boat. The propeller blade has a fixed pitch with diameter of 140 cm, a pitch
ratio of 0.77, an expansion ratio of 0.6 and three blades. Table 1 presents the geometric
parameters of the blade. Figure 2 shows the finite element mesh of the MAU3-60

Table 1. Basic geometric parameters of blade of MAU 3-60 propeller.

r/R C/D P/D hm (8) Zm/D t/D f/D

0.2 0.3108 0.77 9.00 0.0176 0.0430 0.1385


0.3 0.3629 0.77 7.40 0.0265 0.0383 0.1055
0.4 0.4067 0.77 5.50 0.0353 0.0335 0.0824
0.5 0.4406 0.77 3.50 0.0441 0.0288 0.0653
0.6 0.4629 0.77 1.30 0.0529 0.0240 0.0519
0.7 0.4654 0.77 1.08 0.0617 0.0193 0.0414
0.8 0.4340 0.77 3.75 0.0705 0.0146 0.0337
0.9 0.3439 0.77 6.68 0.0794 0.0098 0.0284
1.0 0.0000 0.77 9.10 0.0882 0.0000 0.0000

40

30

20
Y-axis (chordwise, cm)

10

−10

−20

−30

−40
0 10 20 30 40 50 60 70 80
X-axis (spanwise, cm)
Figure 2. Projected view and finite element mesh of MAU3-60 propeller blade.

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
450 H.J. LIN AND J.F. TSAI

propeller blade. The rotational speed n of the propeller is 12 rps. The advancing speed, Va,
of the propeller is 1.68 m/s. A uniform inflow condition is considered. The mesh size of
the propeller blade is 8  16 in the structural analysis and 8  8 in the fluid analysis.
The root is assumed to be fixed ended boundary condition.
The symmetric stacking sequence of graphite/epoxy of T300/1076E was adopted.
Table 2 presents the material properties of MAB (manganese aluminum bronze) and
T300/1076E. The balanced and unbalanced stacking sequences [. . .//90/0]s and
[. . ./2/90/0]s, were considered. The subscript s denotes symmetric with respect to the
middle surface, i.e., camber surface. The first stacking layer with 08 is on the camber
surface. The number of layers is counted from the middle surface to the upper and lower
surfaces of the propeller blades. The number of layers varies with the thickness. For
example, a stacking sequence of [. . .2/90/0]S means 08, 908, , , 08, 908, , , . . . starting
from the camble surface to the suction and pressure sides, respectively. Figure 3 shows the
stacking sequence [. . .//90/0]S in the propeller blade. The generation line (X-axis) of
the propeller blade is taken as the reference of fiber direction of the composites.
The positive fiber orientation is from the root to the leading edge.

Table 2. Material properties of MAB and T300/1076E.

Properties: MAB Data

Longitudinal modulus, E 100 GPa


poisson ratio,  0.3
Density,  8200 Kg/m3
Properties: T300/1076E Data

Longitudinal modulus, E11 145 GPa


Transverse modulus, E22 8.91 GPa
Shear modulus, G12 4.35 GPa
In-plane poisson ratio, 12 0.3
Density,  1577 Kg/m3
Layer thickness 0.001 m

... ...

X
Y [...−q/q/90°/0°/0°/90°/q/q...]
Figure 3. Stacking sequence [. . .//90/0]S of composite in propeller blade.

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
Analysis of Underwater Free Vibrations of a Composite Propeller Blade 451

FINITE ELEMENT METHOD

A geometrically nonlinear degenerate shell element [9] was used. There are five degrees
of freedom at each node, three translations and two rotations. The static finite-element
equation for a geometrically non-linear three-dimensional degenerate shell element is
expressed as [10–13]:

ðKo þ KG Þq ¼ Fext ð5Þ

in which Ko and KG are the linear stiffness matrix and the geometric matrix; q is the nodal
displacement, and Fext is the external forces. The matrices are defined as:
Z
½Ko  ¼ BTo DBo dV ð6Þ

Z  
T
x xy
KG ¼ G G dV: ð7Þ
xy y

In these equations, Bo represents the linear strain-displacement transformation matrix; D


is the material property matrix, and G is a matrix defined by the derivative of the
coordinates.  x,  y and  xy are stresses on the tangent plane of the shell surface.
The centrifugal forces are considered even though the marine propeller rotates relatively
slowly. They may affect the stiffness of the structure and so much be taken into account.
The Lagrange equation and kinetic energy of the propeller blade are used to formulate
the centrifugal force and rotational stiffness [14–16]. The kinetic energy of a propeller
blade is given by:
Z
1
T¼   V2  dV ð8Þ
2
* * *
where  is the density of the material, and V is the velocity (V*2 ¼ V  V ). The
*
velocity of a point includes the translational and rotational speeds, u_ and !, and is
defined as:

* *
* 
* *
V ¼ u_ þ !  X þ u ð9Þ

*  T  T *
_ w_ gT .
*
where X ¼ x, y, z is the position coordinate, ! ¼ !x , !y , !z , and u_ ¼ fu,
_ v,
Equation (9) can be expressed in detail as:
8 9 8 9
> u_ > > !y  ðz þ wÞ  !z  ðy þ vÞ >
>
< = <> > >
=
*
V ¼ v_ þ !z  ðx þ uÞ  !x  ðz þ wÞ : ð10Þ
>
> > > >
>
: > ; >
: ;
w_ !x  ðy þ vÞ  !y  ðx þ uÞ
* *
Computing V 2 and canceling the terms proportional to X  X, which do no
contribute to the Lagrangian, and substituting Equation (10) into Equation (8)

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
452 H.J. LIN AND J.F. TSAI

yields the kinetic energy,


8 9T 8 9 8 9T 8 9
Z>
< u_ >
= >
< u_ >
= Z >
< u_ >
= <u>
> =
1 1
T¼  v_ v_  dV þ  v_ ½A1  v  dV
2 : >
> ; >
: > ; 2 : >
> ; : >
> ;
w_ w_ w_ w
8 9T 8 9 8 9T 8 9
Z >< u>= >
< u>= Z > < x>= <u>
> =
1
þ  v ½A2  v  dV þ  y ½A2  v  dV ð11Þ
2 >
: > ; >
: > ; >
: > ; >
: > ;
w w z w

where A1 and A2 are expressed as:


2 3
0 2!z 2!y
6 2! 0 2!x 7
½A1  ¼ 4 z 5 ð12Þ
2!y 2!x 0
2 3
!2y þ !2z !x  !y !x  !z
6 7
½A2  ¼ 4 !x  !y !2x þ !2z !z  !y 5 ð13Þ
!x  !z !z  !y !2y þ !2x

Applying the finite element method yields the Lagrangian equation of motion,
   
d @T @T
 ¼ Mq€ þ Cq_  KR q  FR ð14Þ
dt @u_ @u

in which q, q_ and q€ are the nodal displacement, the velocity and the acceleration,
respectively; M is the mass matrix; KR is the rotational stiffness matrix; and FR is the
centrifugal force, as shown below.
Z
M¼ ½NT ½N  dV ð15Þ
Z
C¼ ½NT ½A1 ½N  dV ð16Þ
Z
KR ¼ ½NT ½A2 ½N  dV ð17Þ
8 9
Z <x>
> =
FR ¼ ½NT ½A2  y  dV: ð18Þ
: >
> ;
z

Based on the static finite element analysis, the governing equation of motion, neglecting
damping, can be expressed as:

Mq€ þ ðKo þ KG  KR Þq ¼ Fext þ FR : ð19Þ

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
Analysis of Underwater Free Vibrations of a Composite Propeller Blade 453

ADDED MASS FORMULATION

The propeller blade operates in the water. The additional pressure transmitted to the
blade surface, due to the inertia of the water, is represented as added mass. Assume a node
on the propeller blade generating a movement of displacement, the generated velocity can
be transfer to the normal velocity, q_ n , of the node [9], by
q_ n ¼ ½Tq_ ð20Þ
in which [T ] is the transformation matrix between q_ and q_ n .
The vibration of the structure submerged in the water exerts a pressure normal to the
shell surface. The ideal flow of the source is assumed to be distributed on the surface of
the structure. The velocity potential of a source is expressed as [17–19]:
ZZ
1 1
¼  * dS ¼ ½Pf g ð21Þ
4 R *
ðpÞ  RðqÞ
* *
in which  is the source intensity, and RðpÞ and RðqÞ are the position vectors of two points,
p and q. Thus, the normal velocity can be obtained by the dot product of the derivative of
velocity potential and the surface normal vector.
0 1
ZZ
B1 1
q_n ¼ r  n ¼
*
  r@ * C dS ¼ ½C f g:
A
ð22Þ
4 RðpÞ  R
*
ðqÞ

The pressure can be derived using the unsteady Bernoulli equation as


@
p ¼  ð23Þ
@t
where  is the density of the fluid. Substituting Equations (20)–(22) into Equation (23)
€ as,
yields the pressure in terms of q,

p ¼ ½Pf_ g ¼ ½P½C1 q€ n ¼ ½P½C1 ½T q:


€ ð24Þ

Integrating the surface pressure yields the force induced by the fluid
ZZ
FA ¼ p dS ¼ Aq€ ð25Þ

in which A is the so-called added mass matrix. Substituting Equation (25) into
Equation (19) yields the equation of motion in the form,

ðM þ AÞq€ þ ðKo þ KG  KR Þq ¼ Fext þ FR : ð26Þ


For harmonic motion of the structure, the structural displacement q can be expressed
in the harmonic form,

q ¼ qo  ei!t ð27Þ

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
454 H.J. LIN AND J.F. TSAI

in which qo is the amplitude of harmonic motion. According to the free vibration analysis,
Equation (26) can be simplified and rewritten as:


ðKo þ KG  KR Þ  !2 ðM þ AÞ qo ¼ 0: ð28Þ

The solution to the eigenvalue and eigenfunction problem, Equation (28) is determined
using by

Det ðKo þ KG  KR Þ  !2 ðM þ AÞ ¼ 0: ð29Þ

The subspace method was used to solve the eigen-value problem of Equation (29).
However, matrix A in Equation (29) is asymmetric. Therefore, A was replaced with
1/2(A þ AT) [18].

NUMERICAL RESULTS AND DISCUSSION

MAB propeller blade

Manganese–aluminum–bronze (MAB) is commonly used in the manufacture of marine


propeller blades. Table 3 presents its first three natural frequencies and vibration modes.
pffiffiffiffiffiffiffiffiffiffiof the blade in water, !w, is lower than that in air, !a, by 23% to
The natural frequency
30%. Since ! / 1=M, it means that the added mass is almost equal to 0.7 to 1.0 times
of the mass of the blade itself. Thus, the added mass must be considered in evaluating
the dynamic characteristics of the propeller blades. Figures 4 and 5 show the modes
of vibration in air and water. The mode shapes are almost the same in air and water,
although they differ slightly. The first and second modes are pure bending and torsion
modes. The third mode is the secondary bending mode in the direction of span.
Table 3 also presents the effects of rotation on the free vibration characteristics of
marine propeller blades. Marine propeller blades rotate more slowly than fans, jets,
turbines or other types of blade. The rotational speed of the marine blade is 12 rps.
The value of !w when still is almost the same as that when running. Thus, the rotational
effects can be ignored since the marine propeller blade rotates slowly.

Composite propeller blade

Symmetrical blades with balanced and unbalanced stacking sequences were considered
in the analysis. Table 4 shows the first three natural frequencies of the blade with [. . ./152/
90/0]S. The natural frequency of the blade in water is 43–63% lower than that in air.
The difference between the natural frequencies is much larger than that of the blade made

Table 3. Natural frequency (Hz) of propeller blade made of MAB.

MAB MAB
Mode xa xw Mode (xw)still/(xw)running

1 15.2 10.8 1 0.996


2 31.1 24.5 2 0.973
3 36.9 28.6 3 0.991

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
Analysis of Underwater Free Vibrations of a Composite Propeller Blade 455
pffiffiffiffiffiffiffiffiffiffi
of MAB. Since ! / 1=M, then the added mass is almost six times the mass of the
composite blade. The effects of the added mass on the vibration characteristics of the blade
with [. . ./152/90/0]S exceeds that on those of the MAB blade. Figures 6 and 7 plot the
vibration modes in air and water. The mode shapes are almost the same. The first and
third modes are the first pure bending and torsion modes. The second mode is the second

(a) (b)

(c)

Figure 4. Vibration modes of blade made of MAB in air. (a) mode I; (b) mode II and (c) mode III.

(a) (b)

(c)

Figure 5. Vibration modes of blade made of MAB in water. (a) mode I; (b) mode II and (c) mode III.

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
456 H.J. LIN AND J.F. TSAI

Table 4. Natural frequency (Hz) of composite propeller blade.

[. . .152/90/0]S [. . .15/15/90/0/]S
Mode xa xw Mode xa xW

1 23.18 8.71 1 26.82 9.92


2 42.10 22.80 2 41.60 23.06
3 51.86 29.58 3 55.32 29.83

(a) (b)

(c)

Figure 6. Vibration mode of blade with [. . ./152/90/0]S in air. (a) mode I; (b) mode II and (c) mode III.

bending mode in the direction of span. The second and third modes shape of the blade
made of [. . ./152/90/0]S are the same as those of the third and second modes of the MAB
blade, respectively. Also, the contour is shifted toward the trailing edge of the blade with
[. . ./152/90/0]S because the 158 carbon fibers stiffen the leading edge. The anisotropy of the
composite distorts its mode shapes.
Table 4 also presents the first three natural frequencies of the blade with [. . .15/15/
90/0]S. The natural frequency of the blade in water is 50% lower than that in air.
The added mass is also approximately six times that of the mass of the composite blade.
The natural frequencies of the blade with [. . .15/15/90/0]S exceed those of the blade
with [. . ./152/90/0]S, perhaps because the former is less anisotropic.
Table 5 shows the natural frequencies of the blades with [. . ./2/90/0]S, [. . ./2/90/0]S
and [. . .//90/0]S in water. A typical airfoil section of the propeller blade is thicker
at its leading edge than at its trailing edge. The leading edge is stiffer than the trailing
edge. When fibers are negatively stacked, the stiffness of the trailing edge is increased.
Accordingly, negative fiber stacking increases the natural frequency of the blade.
Therefore many of the data obtained for blades with [. . ./2/90/0]S exceed those
obtained for blades with [. . ./2/90/0]S.

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
Analysis of Underwater Free Vibrations of a Composite Propeller Blade 457

(a) (b)

(c)

Figure 7. Vibration mode of blade with [. . ./152/90/0]S in water. (a) mode I; (b) mode II and (c) mode III.

Table 5. Nature frequency (Hz) of composite propeller blade in water.

[. . .152/90/0]S [. . .302/90/0]S [. . .452/90/0]S


Mode !W Mode !W Mode !W

1 8.71 1 8.45 1 8.28


2 22.80 2 23.62 2 23.91
3 29.58 3 29.55 3 27.49
1 10.22 1 9.90 1 9.34
2 20.46 2 23.14 2 24.67
3 32.96 3 33.00 3 31.70
1 9.92 1 10.35 1 9.82
2 23.06 2 25.92 2 26.17
3 29.83 3 30.68 3 29.72

CONCLUSIONS

The free vibration characteristics of metal and composite propeller blades


were analyzed. The effects of rotation and added mass were noted, using the finite
element method. Symmetric composite blades with balanced and unbalanced stacking
sequences were analyzed. The first three natural frequencies and vibration modes are
presented and discussed. Numerical results reveal that the added mass is almost equal to
that of the blade made of MAB, and seven times that of the composite blade. The natural
frequencies of the blade made of MAB and the [. . ./152/90/0]S blade in water are 23–30%
pffiffiffiffiffiffiffiffiffi

and 43–63%, respectively, lower than that of the blade in air. Since ! / 1=M, the
added mass is almost double and six times the masses of the MAB and

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008
458 H.J. LIN AND J.F. TSAI

composite blades, respectively. Submergence in water greatly reduces the natural


frequency of a blade. The mode shapes are almost the same in air and water. The first,
second and third modes are the first bending, torsion and second bending modes in the
direction of the span of the MAB blade, respectively. The second and third modes of the
blade with [. . ./152/90/0]S are the second bending mode in the direction of the span and
torsion mode. The contours of the mode shape are shifted because of the anisotropy of the
composites. A blade with [. . ./2/90/0]S has a higher natural frequency than a blade with
[. . ./2/90/0]S. The rotational effects can be ignored because a marine propeller blade
rotates slowly.

ACKNOWLEDGMENT

The authors would like to thank the National Science Council of the Republic of China
for financially supporting of this research.

REFERENCES

1. Taylor, D. W. (1933). The Speed and Power of Ships, Ransdell, Washington.


2. Cohen, J. W. (1955). On Stress Calculations in Helicoidal Shells and Propeller Blades, Netherlands Research
Center, T. N. O. Shipbuilding and Navigation, Delft Report 21S.
3. Conolly, J. E. (1974). Strength of Propeller, Trans RINA, 103: 139–204.
4. Genalis, P. (1970). Elastic Strength of Propellers – An Analysis by Matrix Methods. Ph.D. Thesis, University
of Michigan, USA.
5. Atkinson, P. (1968). On the Choice of Method for the Calculation of Stress in Marine Propeller, Trans
RINA, 110: 447–463.
6. Ma, J. H. (1974). Stresses in Marine Propellers, Journal of Ship Research, 18: 252–264.
7. Sontvedt, T. (1974). Propeller Blade Stresses – Application of Finite Element Methods, Computers &
Structures, 4(1): 193–204.
8. Atkinson, P. and Glover, E. J. (1988). Propeller Hydroelastic Effects, SNAME on the Propeller’88
Symposium No.21, Jersey, NJ.
9. Lin, H. J. and Lin, J. J. (1996). Nonlinear Hydroelastic Behavior of Propellers using a Finite-element Method
and Lifting Surface Theory, Journal of Marine Science and Technology, 1: 114–124.
10. Bathe, K. J. and Ramm, E. (1975). Finite Element Formulations for Large Deformation Dynamic Analysis,
International Journal for Numerical Methods in Engineering, 9(2): 353–386.
11. Chang, T. Y. and Sawamiphakdi, K. (1980). Large Deformation Analysis of Laminated Shell by Finite
Element Method, Computers & Structures, 13(1): 331–340.
12. Zienkiewicz, O. C. (1991). The Finite Element Method, 4th edn, McGraw-Hill, London.
13. Bathe, K. J. and Bolourchi, S. (1979). Geometrical and Material Nonlinear Plate and Shell Element,
Computers & Structures, 11(2): 23–48.
14. Omprakash, V. and Ramamurti, V. (1989). Dynamic Stress Analysis of Rotating Turbo-machinery
Blade-disk System, Computers & Structures, 32(2): 477–488.
15. Dokainish, M. A. and Rawtani, S. (1971). Vibration Analysis of Rotating Cantilever Plates, International
Journal for Numerical Methods in Engineering, 3(2): 233–248.
16. Bossak, M. A. J. and Zienkiewicz, O. C. (1973). Free Vibration of Initially Stressed Solid with Particular
Reference to Centrifugal Force Effects in Rotating Machinery, Journal of Strain Analysis, 8(4): 245–252.
17. Vorus, W. S. and Hylarides, S. (1982). Hydrodynamic Added-mass Matrix of Vibrating Ship Based on
a Distribution of Hull Surface Source, Trans. SNAME, 89: 397–416.
18. Jennings, A. (1985). Added Mass for Fluid-structure Vibration Problems, International Journal for Numerical
Methods in Fluids, 5(9): 817–830.
19. Hess, J. L. and Smith, A. M. (1964). Calculation of Non-lifting Potential Flow about Arbitrary Three
Dimensional Bodies, Journal of Ship Research, 8(2): 22–44.

Downloaded from http://jrp.sagepub.com at NATIONAL TAIWAN UNIV LIB on October 24, 2008

You might also like