Sie sind auf Seite 1von 10

Free vibration of cantilevered composite plates in air and in water

Matthew R. Kramer
a
, Zhanke Liu
b
, Yin L. Young
a,
a
Department of Naval Architecture and Marine Engineering, University of Michigan, Ann Arbor, MI 48109, USA
b
P.O. Box 3263, Sugar Land, TX 77487
a r t i c l e i n f o
Article history:
Available online 7 August 2012
Keywords:
Added mass
Composite
Hydroelasticity
Cantilevered plate
Fluidstructure interaction
Free vibration
a b s t r a c t
Composite materials are being used more frequently for marine applications due to the advantages of a
higher stiffness- and strength-to-weight ratio, and better corrosion resistance compared to metallic
alloys. Many examples consist of cantilevered structures, such as hydrofoils, propeller and turbine blades,
keels, and rudders. A wide range of analytical and numerical tools exist for the free vibration analysis of
composite structures in air due to their applicability to design problems in the aerospace industry, such
as airplane wings, turbofan and propeller blades, and ight control surfaces. For these aerospace struc-
tures the inertial effects of the uid are typically neglected due to the low relative density of air compared
to the structure. Contrarily, for marine structures, uid inertial (added mass) effects cannot be neglected,
especially for composites with much higher uid-to-solid density ratios. The objective of this work is to
investigate the effects of material anisotropy and added mass on the free vibration response of rectangu-
lar, cantilevered composite plates/beams via combined analytical and numerical modeling. The results
show that the natural frequencies of the composite plate are 5070% lower in water than in air due to
large added mass effects. The added mass is found to vary considerably with material orientation due
to the bend-twist coupling of anisotropic composites, which affects the mode shapes and, consequently,
the uid inertial loads. The analytical method is found to yield accurate results for beam geometries and
offers signicant savings in computational cost compared to the nite element method.
2012 Elsevier Ltd. All rights reserved.
1. Introduction
The use of composite materials for marine applications is
becoming more prevalent due to the advantages of a generally
higher strength- and stiffness-to-weight ratio, and better corrosion
resistance compared to traditional metallic alloys such as steel,
aluminum, or bronze. Additionally, composite materials offer the
ability to elastically tailor the deformation through the design of
the material (e.g. via the laminate stacking sequence), which can
yield better performance over a wider range of operating condi-
tions [9,10,12,1618,21,27].
Many marine applications for which composite materials have
been used consist of cantilevered structures, including propeller
and turbine blades, hydrofoils, keels, and rudders. One important
aspect of the design of these structures is free vibration analysis,
where the natural frequencies and mode shapes are calculated in
order to predict the dynamic structural response, as well as to
identify potential instability limits such as resonance and utter.
Similar design problems may be found in the aerospace industry,
including aircraft wings, turbofan and propeller blades, helicopter
rotors, and wind turbines, and many of these have benetted
greatly from the use of composite materials. Consequently, many
analytical and numerical tools have been developed and validated
for composite structures in air.
Narita and Leissa [19] presented an analytical method for calcu-
lating the dry natural frequencies and mode shapes of very thin
composite plates using the Ritz method. They analyzed a large
number of congurations, varying both the aspect ratio and the
laminate layup sequence, and the results were shown to compare
favorably with both numerical (nite-element) and experimental
results of Crawley [6]. Analytical methods have also been devel-
oped for Timoshenko beams [4,22] for varying geometries and
laminate layup sequences. In each of these studies, the effects of
the uid inertia, i.e. added mass, of the surrounding uid are ne-
glected due to the relatively low density of air compared to the
structural density. This assumption, however, is not typically valid
for marine structures, for which added mass effects are often sig-
nicant. Consequently, methods for the prediction of the free
vibration of composite structures must be developed with consid-
eration for uid inertial effects.
The effect of added mass on the natural frequencies increases as
the ratio of uid-to-solid density increases, and consequently the
added mass of most structures in water may not be neglected,
especially for composite structures, which generally have a lower
0263-8223/$ - see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.compstruct.2012.07.017

Corresponding author.
E-mail addresses: mattkram@umich.edu (M.R. Kramer), zhankeliu2008@gmail.
com (Z. Liu), ylyoung@umich.edu (Y.L. Young).
Composite Structures 95 (2013) 254263
Contents lists available at SciVerse ScienceDirect
Composite Structures
j our nal homepage: www. el sevi er . com/ l ocat e/ compst r uct
effective density than metallic structures. The added mass effects
for cantilevered isotropic plates have been investigated by several
authors, including Marcus [14] and Yadykin et al. [26], to name a
few. In [14], a nite element method is used, where the uid effects
are modeled using inviscid, linearly compressible acoustic uid
elements. The results were found to compare well with the exper-
imental results of Lindholm et al. [11]. In [26], an analytical formu-
lation is presented for a submerged, cantilevered, isotropic plate in
water, where the uid inertial effects are calculated based on po-
tential ow theory. Although several methods for considering uid
inertial effects have been developed, they have previously been ap-
plied to isotropic structures constructed of metallic alloys. Since, in
general, added mass effects are highly dependent on the nature of
the vibrational modes (e.g. bending or twisting), the bend-twist
coupling that results from the material anisotropy present in com-
posite materials introduces more complex added mass effects that
must be quantied.
The objective of this work is to investigate the effects of mate-
rial anisotropy and added mass on the free vibration response of
rectangular, cantilevered composite plates/beams via combined
analytical and numerical studies. An analytical method is rst pre-
sented, followed by numerical modeling via nite element simula-
tions. Convergence and validation studies are shown. The inuence
of added mass and material anisotropy on the free-vibration re-
sponse of composite plates and beams are examined, followed by
a summary of the major ndings.
2. Problem description
The current study is aimed towards cantilevered structures
such as wings, hydrofoils, propeller and turbine blades, keels, and
rudders. Due to the variability in the geometry of these structures,
they are generalized in this study as rectangular, cantilevered
plates and beams for simplicity. It is expected that the major phe-
nomena exhibited by the rectangular plates/beams are representa-
tive of those present for more complicated structures.
In this work, two uid media are considered: air and water, de-
noted as dry and wet congurations, respectively. For the dry
cases, the air is assumed to have a negligible effect and is ignored
(i.e. uid density, q
f
= 0). A schematic for a typical cantilevered
plate or beam is shown in Fig. 1. The dimensions are dened by
its length, width, and thickness, represented by L, b, and t, respec-
tively. The aspect ratio is dened as = L/b, which is in line with
its denition in aerodynamics, and the thickness-to-chord ratio is
dened as t/b. Results for two different geometries will be shown
in this paper: one that is representative of a thin plate, with
= 2.6 and t/b = 3.46%, and one that satises beam assumptions,
with = 12 and t/b = 20%. The dimensions for each geometry are
shown in Table 1. It should be noted that the length of both the
plate and the beam, as well as the material properties, are identi-
cal, and that only the aspect ratio and thickness ratio are varied.
1
A right-handed coordinate system is placed with the xy plane
at the mid-plane of the plate, where the origin is located at the can-
tilevered end and the z-direction is in the thickness direction (see
Fig. 1). The principle axes of the composite material are denoted as
the 12 axes, where the 1-direction is along the effective ber an-
gle, which is oriented at an angle h with respect to the xy axes.
In general, a composite plate/beam is composed of multiple
laminate plies, and each ply may be aligned at a different angle.
It has been shown that, for symmetric layup sequences, an equiv-
alent unidirectional ber angle may be found to represent the
overall load-deformation characteristics, and hence simplify the
coupled uidstructure interaction (FSI) analysis [17,28]. Although
the internal stress distributions of the structure with the equiva-
lent unidirectional ber angle will not match with the actual mul-
ti-layered composite structure, the deformation behavior, and
consequently the natural modes and frequencies, will be equiva-
lent. In this paper, the equivalent ber angle is denoted as h and
is used as a means for parametric studies to explore effect of mate-
rial anisotropy more generally. It should be noted that both the
analytical and numerical methods presented in this paper are
capable of handling generalized composite layups.
The material is modeled as orthotropic, where the 1-direction is
associated with the ber direction, and the stiffness in the 2- and
3-directions is assumed to be equivalent. Under this assumption,
the entire material behavior may be specied by ve properties,
E
1
, E
2
, G
12
, m
12
, and m
23
, where E
i
is the Youngs modulus in the i-
direction and G
ij
and m
ij
are the shear modulus and Poissons ratio
in the ij plane, respectively. Symmetry considerations require
the following: E
2
= E
3
, m
12
= m
13
, G
12
= G
13
and G
23
= E
2
/2(1 + m
23
).
The assumed material properties for the composite plate/beam
are shown in Table 2.
3. Analytical formulation
The analytical solution for the free vibration of composite
beams with consideration for FSI is presented in this section. The
model accounts for the coupling between bending and torsion in-
duced by material anisotropy. The beams center of gravity is as-
sumed to be collocated with the elastic axis to enable
independent study of material coupling effects. In line with Ber-
noulliEuler beam theory, the current model neglects shear defor-
mation, rotary inertia, and warping effects. The added mass
formulas used in the analytical model are based on potential ow
assumptions and strip theory.
Each longitudinal section is restricted to two degrees of free-
dom: z-displacement and x-rotation about the reference axis, de-
noted as h and /, respectively, as shown in Fig. 2. The governing
partial differential equations for such a structural system under
Fig. 1. Diagram of a typical rectangular composite cantilevered plate or beam.
Aspect ratio is dened as = L/b, thickness-to-chord ratio is represented as t/b.
Table 1
Dimensions of the plate and beam considered in this study.
Dimension Plate Beam
L 243.8 mm 243.8 mm
b 92.50 mm 20.32 mm
t 3.200 mm 4.063 mm
= L/b 2.6 12
t/b 3.46% 20%
1
Additional geometries and material congurations are examined in Section 4.4 for
validation studies with published results from the literature. However, all original
results presented in this paper are for one or both of these two geometries.
M.R. Kramer et al. / Composite Structures 95 (2013) 254263 255
free vibration, ignoring both structural and uid damping, have
previously been given by others [24,7,8,13,23,24], and may be ex-
pressed as:
EI
@
4
h
@x
4
K
@
3
/
@x
3
m
@
2
h
@t
2
= 0 (1)
GJ
@
2
/
@x
2
K
@
3
h
@x
3
I
x
@
2
/
@t
2
= 0 (2)
where h = h(x, t) and / = /(x, t) are both a function of the longitudi-
nal location x and time t. Here, EI is the bending stiffness parameter,
GJ is the torsional stiffness parameter, K is the bending-torsion cou-
pling parameter, m is the total mass per unit length, and I
x
is the to-
tal polar mass moment of inertia per unit length about the x-axis.
The total mass and mass moment of inertia include both solid (m
s
and I
x,s
) and uid (m
a
and I
x,a
) components, and are dened as:
m = m
s
m
a
= q
s
bt m
a
(3)
I
x
= I
x;s
I
x;a
= q
s
bt(b
2
t
2
)
12
I
x;a
(4)
where q
s
is the solid density.
The added mass terms are calculated based on potential ow
theory, as described in [5,20]:
m
a
= q
f
p
4
(b
2
cos
2
/ t
2
sin
2
/) ~ q
f
p
4
b
2
(5)
I
x;a
= q
f
p
128
b
4
(6)
where q
f
is the uid density. It should be noted that, although the
added mass contains nonlinear terms in /, these terms may be
approximated based on small angle arguments for frequency
analysis.
The three stiffness parameters (EI, GJ, and K) may be calculated
following the work of Weisshaar and Foist [25] as:
EI = b D
11

D
2
12
D
22
_ _
(7)
GJ = 4b D
66

D
2
26
D
22
_ _
(8)
K = 2b D
16

D
12
D
26
D
22
_ _
(9)
where the expressions of bending stiffness terms D
11
, D
22
, D
66
, D
12
,
D
16
, and D
26
are dependent on the material properties and orienta-
tion, as dened in Appendix A. Notice that the chordwise bending
moment has been neglected because its contribution is assumed
to be small in comparison to the spanwise moment. However, there
is no constraint to the chordwise curvature or camber bending, and
hence it is different from the assumption of chordwise rigidity [25].
Using separation of variables, similar to Banerjee [3] and Wang
et al. [23], it is assumed that h(x, t) = H(x)e
ixt
and /(x, t) = U(x)e
ixt
,
where x is the eigenfrequency. Eqs. (1) and (2) may then be writ-
ten in eigen format:
EIH
(4)
KU
(3)
mx
2
H = 0 (10)
GJU
(2)
KH
(3)
I
x
x
2
U = 0 (11)
where f
(n)
= d
n
f/dx
n
represents the nth spatial derivative of a func-
tion f. By eliminating either H or U from Eqs. (10) and (11), a
sixth-order differential equation is obtained:
W
(6)

I
x
EI x
2
EI GJ K
2
W
(4)

m GJ x
2
EI GJ K
2
W
(2)

mI
x
x
4
EI GJ K
2
W = 0
(12)
where W = H or U. Introducing the non-dimensional length n = x/L
and letting D = d()/dn, Eq. (12) can be re-written as:
(D
6


aD
4


bD
2

bc)W = 0 (13)
where

a =
I
x
EI x
2
L
2
EI GJ K
2
(14)

b =
m GJ x
2
L
4
EI GJ K
2
(15)
c = 1
K
2
EI GJ
(16)
The general solutions of Eq. (13) takes the form [3,23]:
H(n) = A
1
coshan A
2
sinhan A
3
cos bn A
4
sinbn A
5
cos cn A
6
sincn (17)
U(n) = B
1
coshan B
2
sinhan B
3
cos bn B
4
sinbn B
5
cos cn B
6
sincn (18)
where
a = 2(q=3)
1=2
cos(u=3)

a=3
1=2
(19)
b = 2(q=3)
1=2
cos[(p u)=3[

a=3
1=2
(20)
c = 2(q=3)
1=2
cos[(p u)=3[

a=3
1=2
(21)
q =

b

a
2
=3 (22)
u = cos
1
(27

bc 9

b 2

a
3
)=[2(

a
2
3

b)
3=2
[ (23)
The coefcients A
16
and B
16
are related and the relation-
ship may be obtained by substituting Eqs. (17) and (18) into
Eq. (10):
Table 2
Assumed composite and uid properties. It should be noted that, for dry cases, the
uid is neglected (i.e. q
f
= K
f
= 0).
Material Item Symbol Value Unit
Composite Density q
s
1500 kg/m
3
Youngs modulus E
1
171.42 GPa
Youngs modulus E
2
, E
3
9.08 GPa
Shear modulus G
12
, G
13
5.29 GPa
Poissons ratio m
12
, m
13
0.32
Poissons ratio m
23
0.29
Water Density q
f
1000 kg/m
3
Bulk modulus K
f
2.2 GPa
Fig. 2. Description of two analytical degrees of a freedom for a longitudinal strip of
the beam. Note that the center of gravity (CG) and elastic axis (EA) are collocated
and assumed to translate in the z-direction only.
256 M.R. Kramer et al. / Composite Structures 95 (2013) 254263
B
1
= k
a
A
2
=L; B
2
= k
a
A
1
=L (24)
B
3
= k
b
A
4
=L; B
4
= k
b
A
3
=L (25)
B
5
= k
c
A
6
=L; B
6
= k
c
A
5
=L (26)
where
k
a
=
EI a
4
mx
2
L
4
Ka
3
(27)
k
b
=
EI b
4
mx
2
L
4
Kb
3
(28)
k
c
=
EI c
4
mx
2
L
4
Kc
3
(29)
Following Banerjee [3] and Wang et al. [23], the expressions for
the bending rotation H(n), bending moment M(n), shear force S(n),
and torsional moment T(n) may be obtained from Eqs. (17) and
(18):
H(n) =
1
L
dH(n)
dn
=
1
L
(A
1
asinhan A
2
acoshan A
3
bsinbn
A
4
b cos bn A
5
c sincn A
6
c cos cn) (30)
M(n) =
EI
L
2
d
2
H(n)
dn
2
=
EI
L
2
(A
1
a
2
coshan A
2
a
2
sinhan
A
3
b
2
cos bn A
4
b
2
sinbn A
5
c
2
cos cn A
6
c
2
sincn) (31)
S(n) =
1
L
dM(n)
dn
=
EI
L
3
(A
1
a
3
sinhan A
2
a
3
coshan A
3
b
3
sinbn
A
4
b
3
cos bn A
5
c
3
sincn A
6
c
3
cos cn) (32)
T(n) =
GJ
L
dU(n)
dn
=
GJ
L
2
(A
1
ak
a
coshan A
2
ak
a
sinhan
A
3
bk
b
cos bn A
4
bk
b
sinbn A
5
ck
c
cos cn A
6
ck
c
sincn)
(33)
To derive the frequency equation, the following boundary
conditions (at the clamped end n = 0 and at the free end n = 1,
respectively), are applied:
H(0) = H(0) = U(0) = 0 (34)
M(1) = S(1) = T(1) = 0 (35)
By substituting Eqs. (17), (18), (30)(32), and (33) into Eqs. (34)
and (35), the following linear system of equations is obtained:
BA = 0 (36)
where A = [A
1
, A
2
, A
3
, A
4
, A
5
, A
6
]
T
and B takes the following form:
1 0 1 0 1 0
0 a 0 b 0 c
0 k
a
0 k
b
0 k
c
a
2
C
ha
a
2
S
ha
b
2
C
b
b
2
S
b
c
2
C
c
c
2
S
c
a
3
S
ha
a
3
C
ha
b
3
S
b
b
3
C
b
c
3
S
c
c
3
C
c
ak
a
C
ha
ak
a
S
ha
bk
b
C
b
bk
b
S
b
ck
c
C
c
ck
c
S
c
_

_
_

_
where
C
ha
= cosha; C
b
= cos b; C
c
= cos c
S
ha
= sinha; S
b
= sinb; S
c
= sinc (37)
The necessary and sufcient condition for a non-zero solution
to Eq. (36) is D = det[B] = 0, which yields the natural frequencies
of the composite beam.
The frequency, x, may be non-dimensionalized, as in [19]:
X = xL
2

q
s
t
D
o
_
(38)
where D
o
= E
1
t
3
/12(1 m
12
m
21
) corresponds to the reference bend-
ing stiffness at h = 0. All further results for frequency will be non-
dimensionalized accordingly.
4. Numerical formulation
4.1. Governing equations
Numerical simulations have also been performed in order to (1)
validate the results of the analytical model and (2) obtain results
for cases where the beamassumptions are not valid (e.g. cases with
plate-like geometry with low aspect ratio or small thickness-to-
chord ratio, where the chordwise bending moment is not small
compared to the spanwise bending moment), or for cases where
higher-order modes with combined bend, twist, and warping
behavior develop, and the simplied expressions for the added
mass terms as shown in Eqs. (5) and (6) are no longer valid.
The commercial nite element code, Abaqus/Standard [1], has
been used to perform frequency analysis for both wet and dry con-
gurations. For wet congurations, the uidstructure interaction
(FSI) is modeled by coupling an acoustic uid domain to the front
and back faces of the plate via surface-based tie constraints on the
shared boundaries, which enforce the displacement and pressure
compatibility conditions at the uidsolid interface. The other
three wetted faces were not coupled to avoid numerical instability
issues. The effect of neglecting these faces should be negligible,
however, since the front and back plates support the majority of
the pressure forces.
Assuming the uid to be inviscid, irrotational, linearly com-
pressible, and neglecting gravitational effects, the equilibrium
equation is:
@p
@x
q
f

u
f
= 0 (39)
where p is the uid pressure, x is the positional vector, q
f
is the uid
density, and

u
f
is the uid particle acceleration vector. The consti-
tutive behavior of the uid is assumed to obey the following
relation:
p = K
f
@
@x
u
f
(40)
where K
f
is the bulk modulus of the uid and u
f
is the uid particle
displacement vector. The uid domain boundaries are assumed to
be innitely far away, although the computational domain is nite
in size. In order to represent an innite domain on a nite numer-
ical grid, a non-reecting boundary condition is used on all exterior
boundaries, except the wall at the cantilevered end of the plate,
where a fully reective boundary condition is used.
In order to model the structural response, the nonlinear (i.e.
large deformation) solid equation of motion is formulated as an
equilibrium equation (i.e. no external forcing) and the eigenvalues
are extracted via the Lanczos method. The generalized nite ele-
ment equilibrium equation for the structure, which assumes no
structural damping may be written as:
M

u
s
Ku
s
= 0 (41)
where M and K are the structural mass and stiffness matrices,
respectively, and

u
s
and u
s
are the structural acceleration and dis-
placement vectors, respectively. Further details of the formulation
can be found in [1].
4.2. Finite element mesh topology
The nite element mesh consists of two major element sets:
one for the solid domain and one for the uid domain. For dry sim-
ulations, the uid domain is not included in the model. Due to the
M.R. Kramer et al. / Composite Structures 95 (2013) 254263 257
simplicity of the geometry, both domains consist solely of rectan-
gular elements, and as such the domain may be completely speci-
ed by a small number of parameters. The solid domain is
discretized in each dimension using n
i
elements, where i corre-
sponds to the x-, y-, or z-direction, a shown in Fig. 3a. Two different
types of solid elements were tested, including eight-noded solid
elements with incompatible mode (C3D8I), and eight-noded con-
tinuum shell elements with reduced integration (SC8R). Details
of the formulation for each element may be found in [1]. The re-
sults were found to agree favorably for both types of elements.
Hence, for the results shown herein, the shell elements are chosen
for their computational efciency.
The uid domain is generated by extending the solid domain in
all three spatial directions. This is done by specifying the number
of additional layers in each direction, n
f,i
, where i again corresponds
to the x-, y-, and z-directions. For each of these parameters, the ex-
tent of the uid domain may be expressed as a length x
B
, y
B
, and z
B
.
This nomenclature is dened in Fig. 3b, where roughly 1/4 of the
total uid domain is shown. It should be noted that Fig. 3b is
shown for illustrative purposes and is not to scale. The complete
uid domain may be constructed by reecting the portion shown
across both the xy and xz planes. The objective is to place the
boundaries of the uid domain sufciently far away from the plate
in order to eliminate boundary effects. The computational grid
used in the current study is shown in Fig. 4.
The size of the uid element faces with a surface normal pointing
inthe z-directionis made to be identical tothe soliddomainelement
faces and the elements were grown in the z-direction with a growth
rate c = 1.05 (i.e. Dz
i+1
= cDz
i
). The thickness of the rst layer of uid
elements is equal to the thickness of the plate. The uid domain was
modeled using eight-noded acoustic nite elements (AC3D8).
4.3. Convergence studies
As a measure of convergence, the normalized error of the non-
dimensional frequency is chosen, which is dened as
e =
[X
i
(h) X
i;finest
(h)[
X
i;finest
(h)
(42)
where X
i
is the non-dimensional natural frequency corresponding
to the ith mode. A sufcient level of convergence is accepted when
the maximum error for the rst two modes is less than 10
3
. The
convergence study was conducted for the plate with = 2.6 and
t/b = 3.46%, and with a ber orientation angle h = 30. Additional
convergence studies were conducted for other ber orientation an-
gles and aspect ratios with similar results, but the results are not
presented in this paper for the sake of brevity.
In order to determine the required level of solid domain mesh
renement, the three mesh parameters, n
x
, n
y
, and n
z
were system-
atically varied for the dry case. The two parameters n
x
and n
y
were
not chosen independently and instead were chosen to yield ele-
ment faces that were as close to a square as possible. The ratio of
cells is therefore dependent on the aspect ratio of the plate, i.e.
n
x
/n
y
~ . The absolute error of the dry natural frequency for a
selection of grids is shown in Table 3. The nal mesh was chosen
to be [n
x
, n
y
, n
z
] = [72, 27, 7].
Fig. 3. Diagram of nite element mesh topology.
258 M.R. Kramer et al. / Composite Structures 95 (2013) 254263
Once convergence of the solid mesh for dry natural frequency
was demonstrated, the uid mesh parameters were varied in a
similar manner and the absolute error of the wet natural frequency
was calculated for a selection of domain sizes. The results are
shown in Table 4. The mesh was determined to have a sufcient le-
vel of convergence with values of [n
f,x
, n
f,y
, n
f,z
] = [36, 36, 35], which
yields domain extents of [x
B
/L, y
B
/L, z
B
/L] = [0.50, 0.51, 1.24]. The
converged mesh is shown in Fig. 4 for reference.
4.4. Validation studies
Once the converged mesh was identied, the geometry and
material properties were varied in order to compare the dry and
wet frequencies with published results from open literature. In
order to validate the numerical solution of submerged composite
cantilevered plates, two separate validation studies were
performed.
The rst compares the results for a dry composite cantilevered
plate to those of Narita and Leissa [19] in order to ensure that the
solid structural model is able to capture the effects of varying ber
angle and layup sequence. In Fig. 5a, the non-dimensional frequen-
cies for a cantilevered, multidirectional composite plate (layup
sequence [0/30/30/0]) is compared to the analytic results of Narita
and Leissa [19], as well as with numerical and experimental results
of Crawley [6]. The results are found to compare well, especially for
the rst three modes. The current numerical study is then
compared to Narita and Leissas (N & L) analytical model [19] for
unidirectional composite plates with a range of ber angles.
Contours of the magnitude of the relative difference between the
results are shown in Fig. 5b, where it can be seen that the maxi-
mum difference is 0.81%. Similar results were also found for two
other sets of material properties. Hence, the current nite element
model of the dry composite plates is assumed to be sufciently
accurate.
The second necessary validation study relates to the FSI
coupling between the uid and solid domains. Comparisons of
the predicted wet natural frequencies for a cantilevered, steel plate
( = 2, t/b = 1.31%) with the numerical results of Marcus [14] and
experimental results of Lindholm et al. [11] are shown in Fig. 6.
In general, good comparisons are observed. However, the frequen-
cies tend to be slightly overpredicted for all modes compared to
the measured values, particularly for the higher modes. This may
be due to material or uid damping effects, or measurement errors,
particularly for the higher modes.
5. Results and discussion
5.1. Comparison of analytical and numerical methods
The in-air (dry) and in-water (wet) natural frequencies of the
composite plate and beam with varying ber angles, h, are shown
in Fig. 7. The dimensions and material properties of the plate and
beam are shown in Tables 1 and 2, respectively. As shown in
Fig. 7b, both the analytical and numerical results agree well with
each other for the beam geometry. However, although both the
analytical and numerical results showed similar trends for the
plate geometry (see Fig. 7a), the agreement between the two is
not as good because the beam assumptions are violated, and
chordwise deformation and warping are no longer negligible.
The results suggest that the analytical solution is quite accurate
for beams with high and sufciently high thickness-to-chord ra-
tio t/b, and should be used for those cases because of the signicant
savings in computational time. Each nite element computation of
the wetted natural frequencies takes on the order of a few hours of
computing time on four processors (Dual socket six core Intel Core
I7 CPU nodes), but each analytical solution takes only a few sec-
onds on a single processor. These computational savings offer huge
benets, particular for early-stage design studies. Moreover, the
analytical solution allows easy derivation of the characteristic re-
sponse for varying inputs, as well as critical scaling relations for
experimental studies.
5.2. Effects of added mass on natural modes and frequencies
The uid inertial effects are highly dependent on the mode
shapes, which are illustrated in Fig. 8 for the composite plate with
= 2.5 and t/b = 3.46% at select ber angles h. The results show
that the dry and wet mode shapes are very similar, except for small
differences at the higher modes. For 0 < h < 90, the mode shapes
exhibit combined bend and twist deformation starting from the
rst mode because of the anisotropic material properties, and
the twisting deformation increases with h due to reduction of the
effective bending stiffness.
Fig. 4. Converged nite element mesh for wet plate simulations showing the
complete solid domain and roughly one half of the uid domain (Fine mesh:
[n
x
, n
y
, n
z
] = [72, 27, 7], and Large domain: [x
B
/L, y
B
/L, z
B
/L] = [0.50, 0.51, 1.24]).
Table 3
Convergence of dry natural frequency for various structural mesh resolution for the
plate with = 2.6, t/b = 3.46%, and at a ber angle h = 30. The second mode is
chosen since it generally yielded higher levels of error than the rst mode. The ne
mesh was chosen for the remainder of the analyses.
Mesh n
x
n
y
n
z
X
2,dry
e (10
3
)
Coarse 24 9 3 1363.4 8.964
Medium 48 18 5 1364.9 1.273
Fine 72 27 7 1364.6 0.412
Finest 96 36 9 1364.3 0.000
Table 4
Convergence of wet natural frequency for various uid domain sizes for the plate with
= 2.6, t/b = 3.46%, and at a ber angle h = 30. The second mode is chosen since it
generally yielded higher levels of error than the rst mode. The Large domain was
chosen for the remainder of the analyses.
Mesh x
B
/L y
B
/L z
B
/L X
2,wet
e (10
3
)
Small 0.33 0.34 0.17 437.81 41.08
Medium 0.42 0.42 0.46 454.47 4.600
Large 0.50 0.51 1.24 456.29 0.613
Largest 0.58 0.59 2.89 456.57 0.000
M.R. Kramer et al. / Composite Structures 95 (2013) 254263 259
The non-dimensional wet and dry natural frequencies corre-
sponding to the mode shapes seen in Fig. 8 are plotted in Fig. 9
along with the wet-to-dry frequency ratios as a function of ber
orientation angle. It can be seen that in all cases, there is a 50
70% reduction in the natural frequencies due to added mass effects.
The results also show that the wet-to-dry frequency ratios vary
with ber angle and mode number. At an angle of h ~ 20, a cross-
over behavior is observed for the second, third, and fourth modes,
which is due to the change in added mass caused by the switch
from bending dominated to combined bending and twisting mode
shapes.
To better understand the dependence of the added mass (and
hence the in-water frequencies) on the mode shapes, simplied
formulas can be derived using two-dimensional (2-D) potential
ow theory for the added mass and isotropic beam theory for the
resonance frequency. Assuming that the cantilevered beam only
undergoes spanwise bending and twisting deformation with no
Fig. 5. Comparison of the predicted dry natural frequencies of composite plates
obtained using the current nite element model with analytical predictions by
Narita and Leissa [19], as well as with experimental measurements and numerical
predictions by Crawley [6].
Fig. 6. Comparison of current nite element model (FEM) results with numerical
results of Marcus [14] and experimental results of Lindholm et al. [11] for a
submerged, cantilevered, steel plate ( = 2, t/b = 1.31%).
Fig. 7. Comparison of analytical (curves) and numerical (symbols) results for the
rst two modes of (a) composite plates and (b) composite beams with varying ber
angle h when in air (dry) and when fully submerged in water (wet).
260 M.R. Kramer et al. / Composite Structures 95 (2013) 254263
change in the cross sectional geometry, Eqs. (5) and (6) can be used
to estimate the sectional added mass and moment of inertia, which
yields the following simplied equations for the wet-to-dry
frequency ratios for pure bending and pure twisting modes:
Pure bending:
x
wet
x
dry
=

m
s
m
s
m
a
_
= 1
p
4
q
f
q
s
b
t
_ _

1
2
(43)
Pure twisting:
x
wet
x
dry
=

I
x;s
I
x;s
I
x;a

= 1
3p
32
q
f
q
s
b
t
b
2
b
2
t
2
_ _

1
2
(44)
The results of Eqs. (43) and (44) for the plate geometry are plot-
ted as dashed lines in Fig. 9c. Note that the line with the larger
X
wet
/X
dry
value corresponds to the pure twisting mode while the
lower value corresponds to the pure bending mode. As shown in
the gure, the wet-to-dry frequency ratios obtained using these
equations are slightly lower than the numerical values. This is be-
cause Eqs. (5) and (6) assume simple harmonic motion and neglect
shear deformation, rotary inertia, warping and damping effects,
which leads to overestimates of the added mass/inertia. It should
be noted that Eqs. (43) and (44) are not valid for composite
plates/beams except for h = 0 and h = 90 because of coupled
bend-twist deformations, and are not valid for higher modes where
signicant warping occurs. Nevertheless, as shown in Fig. 9c, these
simple equations do provide reasonable estimates of the lower and
upper values, respectively, of the wet-to-dry frequency ratios for
the range of h and modes because the mode shapes are mostly
dominated by combined bending and twisting patterns.
Eqs. (43) and (44) show that the wet-to-dry frequency ratio is
inversely proportional to the uid-to-solid density ratio (q
f
/q
s
).
Hence, as the density of the uid medium increases, the wetted
natural frequency will be reduced. The equations also explain
why, for structures operating in air, inertial effects are often as-
sumed to be negligible because q
f
/q
s
?0. For composite plates,
added mass effects are much more signicant than for steel plates,
since the density of the composite is typically 46 times lower than
steel. Therefore, the wet-to-dry frequency ratios will also be much
lower for composite plates than for steel plates.
Further examination of both the mode shapes and frequency ra-
tios helps to explain the cross-over behavior that is observed at
h ~ 20. In order to do so, it is useful to explore the limiting cases
of h = 0 and h = 90. At h = 0, Modes 1 and 3 are seen to exhibit
pure bending behavior while Modes 2 and 4 exhibit pure twisting
behavior. At h = 90, Modes 1, 2 and 4 are pure bending modes be-
cause of the much reduced effective bending stiffness, and Mode 3
is a pure twisting mode. The values of the wet-to-dry frequency ra-
tios for each of these modes for these two ber angles are relatively
close to the values given by Eqs. (43) and (44) for the associated
deformation pattern. As h increases, the effective EI, GJ, and K
change, which leads to transition of the mode shapes, which in
turn changes the added mass. For Modes 2, 4, and 5, the wet-to-
dry frequency ratios decrease with increasing h because of the in-
crease in added mass caused by greater bending deformations
resulting from the reduction in EI. For Mode 1, the frequency ratio
increases slightly with h because of small decreases in added mass
cause by the increasing twisting deformation near the tip of the
plate. For Mode 3, which is a bending mode at h = 0, the wet-to-
dry frequency ratios increase with h because of the reduction in
added mass caused by the switch to a twisting dominated mode.
Although the 3-D mode shapes are more complicated for compos-
ite plates, the results show that Eqs. (43) and (44) can be use as
simple effective bounds for the wet-to-dry frequency ratios for
all modes and h.
6. Conclusions
This paper has presented results for the free vibration analysis
of a composite cantilevered plate for both in-air and in-water cases
via combined analytical and numerical analysis. The effects of
material anisotropy and uid-to-solid density ratio were examined
in order to quantify the added mass effects for composite plates
and beams vibrating in water.
The analytical model is based on composite beam theory for the
structural response, and strip theory with potential ow assump-
tions for the added mass. Hence, the analytical model is able to ac-
count for the effects of bend-twist coupling induced by material
anisotropy, as well as added inertial resistance provided by the
surrounding uid. The numerical model is based on a nite element
formulation for both the solid and the uid with tie-based displace-
ment constraints at the uidsolid interface. Convergence and vali-
dation studies are shown for the numerical model. The nite
element simulations compared well with published experimental,
analytical, and numerical results for dry composite plates, and for
wet steel plates. The nite element results also compared very well
with analytical solutions of the in-water free vibration of composite
beams with varying ber angles. The comparisons between the -
niteelement andanalytical solutions are not as goodfor the compos-
ite plate with low aspect ratio, and low thickness to chord ratio;
nevertheless, the analytical method is able to capture the general
trend of variation of the in-water resonance frequencies with ber
angle for the composite plate.
For all the cases investigated, the wetted frequencies were
shown to be signicantly lower than the dry frequencies because
of added mass effects, and this effect is more severe for lightweight
composite plates than for heavier, metallic plates. Two simple
formulas, based on two-dimensional potential ow theory, were
Fig. 8. Dry and wet mode shapes for the rst ve modes of the composite plate with
= 2.6, t/b = 3.46%, and varying ber angle h. Contours are colored by the
normalized z-deection.
M.R. Kramer et al. / Composite Structures 95 (2013) 254263 261
derived for the wet-to-dry frequency ratios, and were found to be
useful as simple, effective bounds for the wet-to-dry frequency ra-
tio for all modes and equivalent ber angles for the composite
plate. The results show that the wet-to-dry frequency ratio is in-
versely proportional to the uid-to-solid density ratio, and hence
is responsible for the 5070% reduction in natural frequency when
the composite plates are submerged in water as opposed to in air.
The results also show that the added mass, and hence in-water nat-
ural frequencies, are highly dependent on the mode shapes, which
in turn are highly affected by material anisotropy.
The objective of this paper has been to highlight the importance
of considering added mass effects for composite plates and beams,
where the uid-to-solid density ratio is large, and to quantify these
effects for varying levels of material anisotropy by varying the
equivalent ber angle. While the current study has focused on
the simplied problem of a rectangular plate/beam, the overall ef-
fects are applicable to other submerged cantilevered structures,
such as hydrofoils, appendages, propeller blades, and ns.
The current study has focused on comparisons of the free vibra-
tion response of cantilevered composite plates and beams in air
and fully submerged in water. However, in order to better quantify
and characterize added mass effects for generalized composite mar-
ine structures, varying levels of submergence must be considered.
Motley et al. [15] presents numerical simulations for the free vibra-
tion response of rectangular, cantilevered composite plates for var-
ious plate aspect ratios and ber angles with special focus on the
effects of (1) varying levels of submergence for surface-piercing
(partially-submerged) plates, (2) varying submergence depth for
fully-submerged plates near the free surface, and (3) tip gap effects
for fully-submerged plates near a wall at the free end.
Acknowledgments
The authors are grateful for the nancial support provided by
the Ofce of Naval Research (ONR) and Ms. Kelly Cooper (program
manager) through Grant Nos. N00014-10-1-0170 and N00014-11-
1-0833. Matthew Kramer is supported by the Department of De-
fense (DoD) through the National Defense Science & Engineering
Graduate Fellowship (NDSEG) Program.
Appendix A. Stiffness parameters as a function of ber angle
For a laminate of N layers, the bending stiffness terms are de-
ned as D
ij
=
1
3

N
k=1
(Q
ij
)
k
(z
3
k1
z
3
k
), where (z
k
, z
k1
) denes the
thickness of the kth ply in the z-direction, and

Q
ij
are the trans-
formed reduced in-plane stiffness coefcients for a single lamina:
Q
11
= Q
11
cos
4
h 2(Q
12
2Q
66
) sin
2
h cos
2
h Q
22
sin
4
h
Q
22
= Q
11
sin
4
h 2(Q
12
2Q
66
) sin
2
h cos
2
h Q
22
cos
4
h
Q
66
= (Q
11
Q
22
2Q
12
2Q
66
) sin
2
h cos
2
h Q
66
(sin
4
h cos
4
h)
Q
12
= (Q
11
Q
22
4Q
66
) sin
2
h cos
2
h Q
12
(sin
4
h cos
4
h)
Q
16
= (Q
11
Q
12
2Q
66
) sinh cos
3
h
(Q
12
Q
22
2Q
66
) sin
3
h cos h
Q
26
= (Q
11
Q
12
2Q
66
) sin
3
h cos h
(Q
12
Q
22
2Q
66
) sinh cos
3
h
where h is the ply orientation angle measured positive counter-
clockwise from the x coordinate to the principal material ber coor-
0 15 30 45 60 75 90
0
5
10
15
20
25
30
35
40
45
0 15 30 45 60 75 90
0
5
10
15
20
25
30
35
40
45
0 15 30 45 60 75 90
0.00
0.05
0.10
0.15
0.20
0.25
0.30
0.35
0.40
0.45
0.50
Fig. 9. Comparison of numerical results for non-dimensional dry and wet frequencies for the rst ve modes of the composite plate with = 2.6, t/b = 3.46% for varying ber
angle h. Here, the symbols represent the discrete simulations, and the curves represent a tted spline to better distinguish the curves and show the trends. In Fig. 9c, the
horizontal dashed lines represent the simplied estimations obtained using Eqs. (43) and (44).
262 M.R. Kramer et al. / Composite Structures 95 (2013) 254263
dinate, and Q
ij
are the reduced in-plane stiffness coefcients of indi-
vidual laminae:
Q
11
=
E
1
1 m
12
m
21
; Q
12
=
m
12
E
2
1 m
12
m
21
Q
22
=
E
2
1 m
12
m
21
; Q
66
= G
12
References
[1] ABAQUS. Version 6.11 documentation. 1080 Main Street, Pawtucket, RI 02860;
2005.
[2] Banerjee JR. Explicit frequency equation and mode shapes of a cantilever beam
coupled in bending and torsion. J Sound Vib 1999;224:26781.
[3] Banerjee JR. Explicit analytical expressions for frequency equation and mode
shapes of composite beams. Int J Solids Struct 2001;38:241526.
[4] Banerjee JR. Frequency equation and mode shape formulae for composite
Timoshenko beams. Compos Struct 2001;51:3818.
[5] Brennen CE. A review of added mass and uid inertial forces. Tech. rep. CR
82.010, Port Hueneme (California): Naval Civil Engineering Laboratory;
January 1982.
[6] Crawley EF. The natural modes of graphite/epoxy cantilever plates and shells. J
Compos Mater 1979;13:195205.
[7] Eftekhari M, Mahzoon M, Ziaie Rad S. An evolutionary search technique to
determine natural frequencies and mode shapes of composite Timoshenko
beams. Mech Res Commun 2011;38:2205.
[8] Guo SJ, Banerjee JR, Cheung CW. The effect of laminate lay-up on the utter
speed of composite wings. J Aerosp Eng, Proc Inst Mech Eng, Part G 2003;217:
11522.
[9] Lee Y-J, Lin C-C. Optimized design of composite propeller. Mech Adv Mater
Struct 2004;11(1):1730.
[10] Lin C-C, Lee Y-J. Stacking sequence optimization of laminated composite
structures using genetic algorithm with local improvement. Compos Struct
2004;63(34):33945.
[11] Lindholm US, Kana DD, Chu WH, Abramson HN. Elastic vibration
characteristics of cantilever plates in water. J Ship Res 1965;9(1):1122.
[12] Liu Z, Young Y. Utilization of bend-twist coupling for performance
enhancement of composite marine propellers. J Fluids Struct 2009;25:
110216.
[13] Lottati I. Flutter and divergence aeroelastic characteristics for composite
forward swept cantilevered wing. J Aircraft 1985;22:10017.
[14] Marcus MS. A nite-element method applied to the vibration of submerged
plates. J Ship Res 1978;22(2):949.
[15] Motley MR, Kramer MR, Young YL. Free surface and solid boundary effects on
the free vibration of cantilevered composite plates. Compos Struct, submitted
for publication.
[16] Motley MR, Liu Z, Young YL. Utilization uid-structure interactions to improve
energy efciency of composite marine propellers in spatially varying wake.
Compos Struct 2009;90:30413.
[17] Motley MR, Young YL. Inuence of design tolerance on the hydroelastic
response of self-adaptive marine rotors. Compos Struct 2011;94(1):11420.
[18] Motley MR, Young YL. Performance-based design and analysis of exible
composite propulsors. J Fluids Struct 2011;27:131025.
[19] Narita Y, Leissa A. Frequencies and mode shapes of cantilevered laminar
composite plates. J Sound Vib 1992;154(1):16172.
[20] Newman JN. Marine Hydrodynamics. The MIT Press; 1977.
[21] Soykasap O, Hodges D. Performance enhancement of a composite tilt-rotor
using aeroelastic tailoring. J Aircraft 2000;37(5):8508.
[22] Teh KK, Huang CC. The effects of bre orientation on free vibrations of
composite beams. J Sound Vib 1980;69(2):32737.
[23] Wang K, Inman DJ, Farrar CR. Modeling and analysis of a cracked composite
cantilever beam vibrating in coupled bending and torsion. J Sound Vib
2005;284:2349.
[24] Wang K, Inman DJ, Farrar CR. Crack-induced changes in divergence and utter
of cantilevered composite panels. Struct Health Monit 2005;4:37792.
[25] Weisshaar TA, Foist BL. Vibration tailoring of advanced composite lifting
surfaces. J Aircraft 1985;22:1417.
[26] Yadykin Y, Tenetov V, Levin D. The added mass of a exible plate oscillating in
a uid. J Fluids Struct 2002;17:11523.
[27] Young YL, Baker J, Motley MR. Reliability-based design and optimization of
adaptive marine structures. Compos Struct 2010;92:24453.
[28] Young YL, Liu Z, Motley MR. Inuence of material anisotropy on the
hydroelastic behaviors of composite marine propellers. In: Proceedings of
the 27th symposium on naval hydrodynamics; October 2008.
M.R. Kramer et al. / Composite Structures 95 (2013) 254263 263

Das könnte Ihnen auch gefallen