Sunteți pe pagina 1din 17

Arch Appl Mech (2007) 77: 363–379

DOI 10.1007/s00419-006-0096-7

O R I G I NA L

K. Daneshjou · A. Nouri · R. Talebitooti

Sound transmission through laminated composite


cylindrical shells using analytical model

Received: 6 February 2006 / Accepted: 20 September 2006 / Published online: 21 November 2006
© Springer-Verlag 2006

Abstract Composite structures are often used in aircraft because of advantages offered by a high strength to
weight ratio. Sound transmission through an infinite laminated composite cylindrical shell is studied in the
context of the transmission of airborne sound into aircraft interior. The shell is immersed in an external fluid
medium and contains an internal fluid, and airflow in an external fluid medium moves with a constant velocity.
The different parameters were used to see how laminate specification affected noise transmission. An exact
solution is obtained by solving the vibration equation of laminated composite shell and acoustic wave equations
simultaneously. Transmission losses (TLs) obtained from numerical solution are compared with those of other
authors. The effects of different source condition, structural properties and flight conditions on TL are studied
for a range of values, especially, incident angle of the plane wave, Mach number and flight altitude of aircraft,
stack sequences, angle of warp and damping.
Keywords Vibro-acoustic · Transmission loss · Laminated composite shell · Plane wave

1 Introduction

Laminated composite shells are increasingly being used in various engineering application including aero-
space, mechanical, marine and automotive engineering. With the increase awareness of, and sensitivity to,
structural noise and vibration, research covering the vibration of composite shells has received considerable
attention. Acoustic design and consideration of composite structures involve elastic wave propagation in com-
posite materials and interaction of sound waves with the composite structures. During cruise flight of a modern
aircraft, the vibration of the outer shell of the fuselage is transmitted to the interior and can produce a high
noise level within the cabin. However, the acoustical properties of these light and stiff structures can often be
less than desirable resulting in high aircraft interior noise levels. The development of lightweight structures,
made of composite materials, has lowered the acoustic transmission loss (TL) of such structures and therefore
further increased the acoustic transmission problem.
Noise transmission, measured by TL through the circular shell, has been studied by many authors. Smith
[20] presented a theoretical study of transmission of sound energy through a thin, isotropic elastic cylindrical
shell from an oblique plane wave excitation. White [22] investigated sound transmission into finite cylindrical
shell and found two important characteristics, the ring and coincidence frequencies, at which TL takes on min-
ima. Koval [9,10] extended Smith’s work to present an analytical model for predicting of TL for isotropic and
orthotropic shells. Blaise et al. [2,3] then extended Koval’s work to orthotropic and multi-layered orthotropic
shells, considering the shells excited by an oblique plane sound wave.
K. Daneshjou (B) · A. Nouri · R. Talebitooti
Department of Mechanical Engineering, Iran University of Science and Technology, Tehran, Iran
E-mail: kdaneshjo@iust.ac.ir
E-mail: ali_nor@mail.iust.ac.ir
E-mail: rt_talebi@mail.iust.ac.ir
Tel.: +98-21-77452169
364 K. Daneshjou et al.

The theoretical study of Koval [11], for an infinite cylindrical shell, provided the first model for noise
transmission loss of composite constructions. Koval’s mathematic model was based on the shell modal imped-
ance, and showed more noise can be transmitted through laminated fiber-reinforced structures than through
isotropic structures in the high frequency range. Roussos et al. [17] gave a report made in the NASA Langley
Research Center about the theoretical and experimental study of noise transmission through composite plates.
Tang et al. [21] considered an infinite cylindrical sandwich shell excited by an oblique plane sound wave.
Analytical and experimental studies are conducted by Lee and Kim [12] to understand the characteristics of
sound transmission through the infinitely long circular cylindrical shell. In most of analytical studies surveyed
above, a simplified shell theory, typically bending approximations or Donnell–Mushtari equation, were used
to model the shell motion. In most of literature survived above, except Lee and Kim [12] only the equation of
motion in the transverse direction was used to describe the shell motion and in-plane equations were completely
neglected, which may not be valid in general. In some cases, numbers of terms used in the series solution were
apparently insufficient to provide converged solutions, which could have affected the very large TLs estimated
results.
In this paper, analytical model of sound transmission through an infinite laminated composite cylindrical
shell excited by an incident oblique plane sound wave is investigated. An aircraft in flight with an external
airflow is modeled as an infinite cylindrical shell. An exact solution was obtained in a series form using clas-
sical laminated shell vibration, without ignoring any of three directions of shell. The shell is assumed to be
immersed in a fluid media and excited by an incident oblique plane sound wave. The properties of the inter-
nal and external fluids surrounding the shell may be different. Acoustic wave equations and shell equations,
which are coupled with one another, lead to vibro-acoustic equations. To make sure an enough number of
modes are included in the analysis, the convergence checking is preformed. In addition, comparison of TL is
made between an aluminum shell, and a composite shell with laminated graphite/epoxy. Finally, the effects of
different source conditions, structural properties and flight conditions on TL are studied for a range of values,
especially, incident angle of the plane wave, Mach number and flight altitude of aircraft, stacking sequence,
angle of warp and damping.

2 Model specification

The specific problem consists of an oblique plane sound wave with angle γ impinging upon a flexible laminated
cylindrical composite shell with mid surface radius R, and included the reflection and scattering of the incident
wave, and the effect of an external airflow (see Fig. 1). It is assumed here the interior cavity inside the shell
is totally absorbing. Therefore only an inward-traveling wave exists in the shell interior. Although this is not
representative of an aircraft interior, the model allows the transmission loss of cylindrical wall to be studied in
a straightforward manner. In the analysis, all waves will be assumed to have the same dependence on the axial
co-ordinate z and the cylinder will be assumed infinitely long. The fluid media in the external and the internal
space are defined by the density and the speed of sound: {ρ1 , c1 } and {ρ3 , c3 } respectively.

3 Governing equations of the fluids

Due to the existence of airflow in the external fluid medium, the external pressure, which is the summation of
the incident wave p I and the reflected wave p1R , satisfies the following wave equation [5,6]:

 2
  ∂
c12 ∇ 2 p I + plR − +V ·∇ ( p I + plR ) = 0 (1)
∂t

where ∇ 2 is the Laplacian operator in the cylindrical coordinate system. The internal pressure of cavity satisfies
the acoustic wave equation:

∂ 2 p3T
c32 ∇ 2 p3T − =0 (2)
∂t 2
Sound transmission through laminated composite cylindrical shells using analytical model 365

Fig. 1 Schematic diagram of the single cylindrical shell: 2-D model

4 Governing equations of the shell

For a composite shell, according to Kirchhoff hypothesis of neglecting shear deformation and the assumption
that the εz is negligible, the stress–strain equation for an orthotropic layer may be written as [16,18]
⎡ ⎤ ⎡ ⎤⎡ ⎤
σ1 Q 11 Q 12 0 ε1
⎣ σ1 ⎦ = ⎣ Q 21 Q 22 0 ⎦⎣ ε2 ⎦ (3)
σ12 0 0 Q 66 ε12

It should be noted that we described the fiber coordinates of orthotropic, as 1 and 2, where direction 1 is parallel
to the fibers and 2 is perpendicular to them. The material constants Q i j are defined in terms of the material
properties of the orthotropic ply:

1 1
Q 11 = E 1 , Q 22 = E 2
  (4)
ν21 ν12
Q 66 = G 12 , Q 12 = E1 = E2 ,  = 1 − ν12 ν21
 
where E 1 and E 2 are module of elasticity in the directions 1 and 2, respectively; G 12 is module of shear
stiffness and νi j (i, j = 1, 2, i  = j) are Poisson’s ratios.
The orientation of the fiber makes the angle θk with the axis of z (Fig. 1). The transformation of stresses
from coordinates 1, 2 to the y, z coordinates can be preformed by using the transformation matrix. This
transformation matrix is
⎡ ⎤
cos2 θk sin2 θk 2 cos θk sin θk
[T ] = ⎣ sin2 θk cos2 θk −2 cos θk sin θk ⎦ (5)
− cos θk sin θk cos θk sin θk cos2 θk − sin2 θk

The transformed stiffness constants Q̄ i j of the kth layer can be written as

[ Q̄] = [T ]−1 [Q][T ] (6)


366 K. Daneshjou et al.

The stress–strain relationship for an element of material in the kth lamia is obtained as below:
⎡ ⎤ ⎡ ⎤ ⎡ ⎤
σz Q̄ 11 Q̄ 12 Q̄ 16 εz
⎣ σ y ⎦ = ⎣ Q̄ 21 Q 22 Q̄ 26 ⎦ ⎣ ε y ⎦ (7)
σ yz k Q̄ 61 Q̄ 62 Q̄ 66 ε yz k
k

The extensional, bending-extensional coupling and bending stiffness Ai j , Bi j and Di j for laminate are as
follows:

N
Ai j = Q ikj (ξk − ξk−1 )
k=1

1 N
 
Bi j = Q ikj ξk2 − ξk−1
2
(8)
2
k=1

1 N
 
Di j = Q ikj ξk3 − ξk−1
3
3
k=1

where ξ is distance from mid-surface. Replacing y = Rϕ, in case of cylindrical shell, the following mid-surface
strain and curvature changes are obtained:
 
∂u 1 ∂v ∂v 1 ∂u
ε0z = , ε0ϕ = + w , γ0zϕ = +
∂z R ∂ϕ ∂z R ∂ϕ (9)
 
∂ 2w 1 ∂v ∂ 2w ∂
v 2 ∂ 2w
k z = − 2 , kϕ = 2 − , k zϕ = −
∂z R ∂ϕ ∂ϕ 2 ∂z R R ∂ϕ∂z
where {u, ν, w} are the displacements of the shell at the neutral surface in the axial, circumferential and radial
directions respectively. The forces N and moments M resultant, obtained by integrating the stresses over the
shell thickness, are
⎡ ⎤ ⎡ ⎤⎡ ⎤
Nz A11 A12 A16 B11 B12 B16 ε0z
⎢ Nϕ ⎥ ⎢ A12 A22 A26 B12 B22 B26 ⎥⎢ ε0ϕ ⎥
⎢ ⎥ ⎢ ⎥⎢ ⎥
⎢ Nzϕ ⎥ ⎢ A16 A26 A66 B16 B26 B66 ⎥⎢ γ0zϕ ⎥
⎢M ⎥= ⎢B ⎥ ⎢ ⎥ (10)
⎢ z ⎥ ⎢ 11 B12 B16 D11 D12 D16 ⎥⎢ k z ⎥
⎣M ⎦ ⎣B ⎦ ⎣ ⎦
ϕ 12 B22 B26 D12 D22 D26 kϕ
Mzϕ B16 B26 B66 D16 D26 D66 k zϕ
Equations of motion of a laminated composite thin cylindrical shell in cylindrical coordinate can be
expressed by [13,16,18]
∂ Nz 1 ∂ Nzϕ ∂ 2u
+ + qz = − I¯ 2 (11)
∂z R ∂z∂ϕ ∂t
 
1 ∂ Nϕ ∂ Nzϕ 1 1 ∂ Mϕ ∂ Mzϕ ∂ 2v
+ + + + qϕ = − I¯ 2 (12)
R ∂ϕ ∂z R R ∂ϕ ∂z ∂t
 
Nϕ ∂ Nz
2 2 ∂ Mzϕ
2 1 1 ∂ Mϕ ∂ Mzϕ ∂ 2w
− + + + + + qr = − I¯ 2 (13)
R ∂z 2 R ∂z∂ϕ R R ∂ϕ ∂z ∂t
In above equations qz , qϕ and qr are external forces (per unit area) in the axial, circumferential and radial
directions, respectively, and t represents the time variable, and
 
I2
I¯ = I1 + (14)
R
N h k

Ii = ρ k ξ i−1 dξ , i = 1, 2 (15)
k=1h
k−1
Sound transmission through laminated composite cylindrical shells using analytical model 367

where Ii is the mass inertia term, ρ k is the mass density of kth layer of the shell per unit mid-surface area,
h k−1 and h k are distances from the reference surface to the kth layer (see Fig. 1) and N is the number of layers.
The equations of motion can be written in terms of displacements
L i j u i + Mi j ü i j = q (16)
where
u i = [u, v, w]T (17)

− I¯ i = j
Mi j = (18)
0 i = j
and
   T
q = 0, 0, p1I + p1R − p3T (19)
In addition, the L i j coefficients are given below:

∂2 A16 ∂ 2 A66 ∂ 2
L 11 = A11 + 2 +
∂z 2 R ∂z∂ϕ R 2 ∂ϕ 2
∂ 2 Ā12 ∂ 2 Ā26 ∂ 2
L 12 = Ā16 2 + + 2
∂z R ∂z∂ϕ R ∂ϕ 2

∂ A ∂ ∂3 B16 ∂ 3 B17 ∂ 3 B26 ∂ 3
L 13 = A∗12 + 26 − B11 3 − 3 − −
∂z R ∂ϕ ∂z R ∂z 2 ∂ϕ R 2 ∂z∂ϕ 2 R 3 ∂ϕ 3
 
∂2 A ∂2 A ∂2
L 22 = A66 2 + 2 26 + 22
∂z R ∂z∂ϕ R 2 ∂ϕ 2 (20)
∂ Ā∗22∂ ∂3 B̄17 ∂3 B̄26 ∂3 ∂3
B̄22
L 23 = Ā∗26 + − B̄16 3 − −3 2 − 3
∂z R ∂ϕ ∂z R ∂z ∂ϕ2 R ∂z∂ϕ 2 R ∂ϕ 3
∗ ∗ ∗
∗ ∂ 4B26 ∂ B22 ∂ ∂ D16 ∂ 4
A 2 2 2 4
L 33 = 22 − 2B12 − − 2 + D 11 + 4
R ∂z 2 R ∂z∂ϕ R 2 ∂ϕ 2 ∂z 4 R ∂z 3 ∂ϕ
D17 ∂ 4 D26 ∂ 4 D17 ∂ 4
+2 2 2 2 + 4 3
R ∂z ∂ϕ R ∂z∂ϕ 3 R 4 ∂ϕ 4
L i j = L ji
B12 D66
Ā12 = A12 + + A66 +
R R
Bi j
Āi j = Ai j + (i j = 16, 22, 26, 66)
R
B17 = B12 + 2B66
D17 = D12 + 2D66 (21)
Di j
B̄i j = Bi j + (i j = 16, 17, 22, 26, 66)
R
B̄i j
Ai j = Āi j + (i j = 22, 26, 66)
R

A B 
Ai∗j , Bi∗j =
ij ij
,
R R
 

Āi j B̄i j
∗ ∗
Āi j , B̄i j = ,
R R

Equation (16) is the shell equation of motion, respectively, in the axial, radial and circumferential direc-
tions. Also, the coupling effect between the shell and acoustic medium is represented by the pressure term in
this equation.
368 K. Daneshjou et al.

5 Boundary conditions at the fluid–structure interfaces

On the internal and external shell surfaces, the particle velocities of the acoustic media in the normal direction
have to be equal to the normal velocity of the shell. These results are shown in the following equations:
  2
∂( p I + p1R )  ∂
 = −ρ1 +V ·∇ w (22)
∂r  ∂t
r =R

∂ p3T  ∂ 2w
 = −ρ3 (23)
∂r  ∂t 2
r =R

6 Solution of the equations

The harmonic incident plane wave p I in cylindrical geometry shown in Fig. 1 can be expressed as [3,12,14]



p I (r, z, ϕ, t) = P0 εn (−j)n Jn (k1r r ) exp [j(ωt − k1z z − nϕ)] (24)
n=0

where εn is the Neumann factor given by



1 (n = 0)
εn = (25)
2 (n ≥ 1)

and

k1z = k1 cos γ , k1r = k1 sin γ (26)

where k1 is the wave number in the moving medium and Jn is the Bessel√function of the first kind of integer
order n, n = 0, 1, 2, 3, . . . , P0 is the amplitude of the incident wave, j = −1 and ω is the angular frequency.
This incidence wave propagates, in the moving medium, according to the convected wave equation
 2  
∂ ∂ ∂2 p ∂2 p
+V p= c12 + 2 (27)
∂t ∂z ∂x2 ∂z

Substitution of Eq. (24) into (27) and x = r cos ϕ yields


 
ω 1
k1 = (28)
c1 1 + M1 cos γ

where, M1 = (V c1 ) is the Mach number of the external flow.
The waves radiated from the shell to the outside and into the cavity, p1R and p3T , can be represented as



p1R (r, z, ϕ, t) = R 2
P1n Hn (k1r r ) exp [j(ωt − k1z z − nϕ)] (29)
n=0

p3T (r, z, ϕ, t) = T 1
P3n Hn (k3r r ) exp [j(ωt − k3z z − nϕ)] (30)
n=0

According to the hypothesis of an internal non-resonant medium, sound pressure in cavity medium, is true
except for r = 0 [4,8].
Sound transmission through laminated composite cylindrical shells using analytical model 369

where, Hn1 and Hn2 are the Hankel functions of the first and second kind of integer order n, respectively.
The former represents the incoming wave and the second the outgoing wave. Three components of the shell
displacements can be expressed as [19,21]


w(z, ϕ, t) = Wn exp [j(ωt − k1z z − nϕ)] (31)
n=0


u(z, ϕ, t) = j Un exp [j(ωt − k1z z − nϕ)] (32)
n=0

v(z, ϕ, t) = j Vn exp [j(ωt − k1z z − nϕ)] (33)
n=0

Because the traveling waves in the acoustic media and inside the shell are driven by the incident-traveling
wave, the wave numbers (or trace velocities) in the z direction should match throughout the system, therefore
k3z = k1z . Then, the following equation can be obtained as

k3r = k32 − k1z 2 , k = ω (34)
3
c3
Substituting Eqs. (24) and (29)–(33) into the shell equation [Eq. (16)] and two boundary conditions [Eqs. (22),
(23)] results in five equations of motion. These five equations involve with six variables: the amplitudes of
the outgoing and incoming waves in the exterior cavity, the transmitted wave in the interior cavity, and three
displacements of the shell structure. Therefore, the solutions can be obtained as the ratios to the one of the
variables, the pressure amplitude of the incoming wave in this case. The ratio of the amplitudes of the input
and transmitted waves obtained this way allows the transmission loss to be obtained. Therefore, the following
five equations are obtained for each circumferential mode number:
   
2 A16 A66 2 ¯ 2 Ā12 Ā26
A11 k1z +
2
nk1z + 2 n − I ω Un + Ā16 k1z + 2
nk1z + 2 Vn
R R R R
 ∗∗ 
A 3B B B
+ A∗12 k1z + 26 n + B11 k1z
16 2 17 26
3
+ k n + 2 k1z n 2 + 3 n 3 Wn = 0 (35)
R R 1z R R
   
Ā12 Ā26 2 A26 A 3
2
Ā16 k1z + nk1z + 2 Un + A66 k1z 2
+ k1z n + 22 n − ¯ω2 Vn
I
R R R R2
 
∗ Ā∗22 B̄17 2 3 B̄26 B̄22 3
+ Ā26 k1z + n + B̄16 k1z +
3
k n + 2 k1z n + 3 n Wn = 0
2
(36)
R R 1z R R
 
A∗ 3B16 2 B17 B26
A∗12 k1z + 26 n + B11 k1z 3
+ k1z n + 2 k1z n 2 + 3 n 3 Un
R R R R
 

Ā22
∗ B̄17 2 3 B̄26 B̄22 3
+ Ā26 k1z + n + B̄16 k1r +
3
k n + 2 k1z n + 3 n Vn2
R R 1z R R
⎧ ∗ ∗ ∗ ⎫

⎪ Ā22 + 2B ∗ k 2 + 4 B̄26 nk1z + 2 B̄22 n 2 + D11 k 4 + 4 D16 k 3 n ⎪ ⎪
⎨ 12 1z 1z ⎬
R R R2 R 1z
+ Wn

⎪ 2D 4D D ⎪

⎩+ 17
k n +
2 2 26
n k1z + 4 n − I¯ω
3 22 4 2 ⎭
R 2 1z R3 R
  R  1  T
− Hn2 (k1r R) P1n + Hn (k3r R) P2n = P0 εn (− j)n Jn (k1r R) (37)

R 2
P1n Hn (k1r R)k1r − ρ1 (ω2 + v 2 k1z
2
+ 2vωk1z )Wn = −P0 εn (− j)n Jn (k1r R)k1r (38)

R 1
P3n Hn (k3r R)k3r − ρ3 ω2 Wn = 0 (39)
R , P T , U , V and W are obtained in terms of P by solving this five
Five unknown coefficients P1n 3n n n n 0
equations.
370 K. Daneshjou et al.

7 Transmission loss

The transmission loss (TL) of the shell is defined as the ratio of the transmitted power W T and the incident
power W I per unit length of the cylinder:

WI
TL = 10 log10 (40)
WT

where

cos(γ )P02
WI = ×R (41)
ρ1 c1
⎧ 2π ⎫
1 ⎨  ⎬
W T = Re p3T .∂ ∂t (w)∗ Rdϕ , r = R (42)
2 ⎩ ⎭
0

where Re{·} and the superscript ∗ represent the real part and the complex conjugate of the argument, respec-
tively. Substitution of Eqs. (30) and (31) for p3T , w into Eq. (42) yields an expression for the components of
WnT :

2π
1 # T 1 $
WnT = Re P3n Hn (k3r r ) · ( jωWn )∗ cos2 [nϕ]Rdϕ where r = R
2
0
πR # T 1 $
= × Re P3n Hn (k3r R) · ( jωWn )∗ (43)
εn

where


WT = WnT (44)
n=0

Finally, the transmission loss can be obtained by substituting Eqs. (41) and (44) into (40).

∞ # T $
Re P3n × Hn1 (k3r R) × ( jωWn )∗ × ρ1 c1 π
TL = −10 log10 (45)
n=0
εn cos(γ )P02

The average power transmission coefficient τ̄ is given as [15]

γmax

τ̄ = τ (γ ) sin γ cos γ dγ (46)


γmin

where,τ (γ ) is the power transmission coefficient calculated for the incident angle γ , in addition γmax and γmin
are the critical angles of incidence upon the shell [9]. In other words, incidence wave out of this range, is totally
scattered with no transmission into the shell interior. The integration in Eq. (46) is done numerically by the
Simpson’s rule using an integration step-size of 2◦ [7]. The average TLav shown in Fig. 2 is given as [1]
 
1
TLav = 10 log10 (47)
τ̄
Sound transmission through laminated composite cylindrical shells using analytical model 371

80

70

60

50

TL (dB)
40

30

20

10
101 102 103 104
Frequency (Hz)
Fig. 2 Transmission loss (TL) averaged for random incident angles

Mode Number Calculation of TL at each


n=n+1 Frequency

No

ABS (TL(n+1)-TL(n))<10–7

Yes

Find Optimum Mode Number

Fig. 3 Algorithm for identifying the optimum mode number

8 Convergence algorithm

Equations (24) and (29)–(31) are obtained in series form. Therefore, enough numbers of modes should be
included in the analysis to make the solution converge. When insufficient number of modes is used in the
calculation, the resulting TL becomes overestimated. Very high TLs for a relatively thin shell reported in the
work by Tang et al. [21] are apparently caused by such a non-converged solution. Once the solution converges
at a given frequency, it can be assumed to converge in all frequencies lower than that, because more terms are
necessary to be used in the calculation for a higher frequency. Therefore, an iterative procedure is constructed
in each frequency, considering the maximum iteration number. Unless the convergence condition is met, it
iterates again. When the TLs calculated at two successive calculations are within a pre-set error bound, the
solution is considered to have converged. Figure 3 shows the concept of this convergence. Changes in the
calculated TL as the number of modes increases are shown in Fig. 4 for the case of a composite shell specified
in Table 1 driven at 1,000 Hz. Figure 5 shows the convergence trend for the same case but at 10,000 Hz, which
indicates that with increasing the frequency, the number of modes for convergence is increased.

9 Validation

Figure 6 gives good validity to our present method. A ten-layered composite shell using “modal
impedance method” offered by “Koval” [11] and the present study were compared. The plies were arranged in a
[0◦ , 90◦ , 45◦ , −45◦ , 0◦ ] S pattern. Comparing these two methods indicates good agreements especially in high
372 K. Daneshjou et al.

45

40

35

TL (dB)
30

25

20

15
0 5 10 15 20 25 30 35
Mode Number
Fig. 4 Mode convergence diagram for the laminated composite shell at 1,000 Hz

Table 1 Structural and environmental properties

Shell Cavity Ambient


Material (fluid) Al Graphite/epoxy Air Air
Density (kg/m3 ) 2,760 1,600 1.29 1.29
E z (GPa) 72 137.9 – –
E θ (GPa) 72 8.96 – –
E zθ (GPa) 27.7 7.1
νzθ 0.3 0.3 – –
Sound speed (m/s) – – 340 340
Incidence angle (degrees) 45

60

55

50
TL (dB)

45

40

35
0 50 100 150 200 250 300
Mode Number
Fig. 5 Mode convergence diagram for the laminated composite shell at 10,000 Hz

frequency. But in low frequency, a significant difference revealed because inertial terms in longitudinal and
circumferential directions were neglected in his work. In addition, tabular data is provided to give detailed
results (Table 2). The basic shell specification and environmental conditions used in the study are listed in
Table 1.
Sound transmission through laminated composite cylindrical shells using analytical model 373

100
Present study
Koval
80

60

TL (dB)
40

20

-20

-40
100 101 102 103 104
Frequency (Hz)
Fig. 6 Comparison of present study with Koval’s for ten-layer composite shell

Table 2 Comparison of present study with Koval method

Frequency (Hz)
10 100 1,000 5,000 10,000 20,000 30,000 40,000
Present method 13.04 16.41 20.17 29.03 18.5 59 70.8 78.62
Koval method 25.63 15.1 17.42 26.1 14.86 56 67.8 75.62

100
γ=30
γ=45
80 γ=60

60

40
TL (dB)

20

-20

-40
100 101 102 103 104
Frequency (Hz)
Fig. 7 TL curves for the ten-layered composite shell with respect to γ incidence angle

10 Numerical results

Numerical results have been generated for typical geometry of a narrow-bodied jet fuselage made of the lami-
nated composite cylindrical shell with radius R = 1.83 m and total thickness h = 1.59 mm. Both internal and
external fluids are considered at sea level conditions and external flow Mach number and incident angle are
assumed as M1 = 0, γ = 45◦ for the purpose of this study except where noted. Each layer of the composite
shell is made of graphite/epoxy (Table 1). The plies were arranged in a [0◦ , 90◦ , 45◦ , −45◦ , 0◦ ] S pattern. For
reference purpose, the same radius and thickness aluminum shell is considered. Parametric numerical studies
of transmission loss (TL) are conducted for broadband frequency. These studies provide insight into the effect
of the acoustic properties of the fluids and the structural or material parameters of the shells on TL.
374 K. Daneshjou et al.

Table 3 Flight conditions

Name
First condition Second condition Third condition
Altitude (m) 3,050 7,600 10,650
Density (kg/m3 ) 0.9041 0.5489 0.3795
Sound speed (m/s) 328.558 309.966 296.556

120
1st Condition
100 2nd Condition
3nd Condition

80

60
TL (dB)

40

20

-20

-40
100 101 102 103 104
Frequency (Hz)
Fig. 8 TL curves for three ambient conditions

The effects of incident angle are shown in Fig. 7 at γ = 30◦ , 45◦ , 60◦ . Inspection of this figure shows
increasing of γ tends to enhance the TL of the shell in stiffness-controlled region (below the ring frequency)
and the coincidence frequency is shifted downwards.
Different ambient conditions are considered here (Table 3). A higher flight altitude affecting density and
sound velocity of fluid will lead to a larger acoustic impedance mismatch between the fluids inside and outside
the shell. As shown in Fig. 8, increasing of acoustic mismatch makes the TL increases in broadband of fre-
quency. For instance in some region it can be enhanced more than 20 dB.
External airflow can influence the axial and radial wave number. Fig. 9, shows the effect of Mach numbers
with M = 0, 0.2, and 0.4, on TL. With increasing Mach number, the TL is descending in stiffness-con-
trolled region, whereas in mass-controlled region (between ring and coincidence frequencies) is ascending.
The coincidence frequency is shifted upwards with increase of Mach number.
Figure 10 shows a comparison between the transmission loss for a ten-layered composite shell and an
aluminum shell with the same radius and thickness. Since, the composite shell is stiffer than the aluminum
shell, its TL is upper than that of aluminum shell in the stiffness-controlled region. However, as a result of
lower density of composite shell, it does not appear to be effective as an aluminum shell in mass-controlled
region. Due to the fact that the critical frequency of a composite shell is lower than it is for an aluminum shell,
so that the TL curve is never able to reach a full mass law behavior because the ring frequency and the critical
frequency are closer together. In the frequencies upper than coincidence frequency, the trend is too similar as
stiffness-controlled region.
The effect of angle of warp θk of the composite shell on TL is shown in Fig. 11. In this Figure, all of
the layers have the same angle θk . The curves in this figure correspond to θk = 0◦ , 30◦ , 45◦ , 60◦ , 90◦ and
the incidence angle of the plane wave is γ = 30◦ . As illustrated, noise transmission of a composite shell is
sensitive to the angle of warp and suggests the possibility of tailoring a composite shell a specific need. This
might be the one of the advantages of a composite shell over an aluminum shell, as regards noise attenuation.
Figure 12 shows the effect of the composite material on TL. Materials chosen for the comparison are graph-
ite/epoxy, glass/epoxy and aramid/epoxy (Table 4). The figure shows that material must be chosen properly to
enhance TL at stiffness-control zone. The best results are obtained for graphite/epoxy shell, which represents a
desirable level of TL at stiffness-control zone. It is readily seen that, in higher frequency, as a result of density
Sound transmission through laminated composite cylindrical shells using analytical model 375

100
M=0
M=0.2
80
M=0.4

60

TL (dB)
40

20

-20

-40
100 101 102 103 104
Frequency (Hz)
Fig. 9 TL curves for the ten-layered laminated shell with respect to Mach number

120
10-layers-Graphite/Epoxy
100 Aluminum

80

60
TL (dB)

40

20

-20

-40
100 101 102 103 104
Frequency (Hz)
Fig. 10 Comparison between Aluminum and ten-layer laminated composite shell

100
θ=0
80 θ=30
θ=45
θ=60
60 θ=90

40
TL (dB)

20

-20

-40
100 101 102 103 104
Frequency (Hz)
Fig. 11 Effect of fiber orientation on TL, at γ = 30◦ for a ten-layered composite shell
376 K. Daneshjou et al.

100
Graphite/Epoxy
Aramid/Epoxy
80
Glass/ Epoxy

60

TL (dB)
40

20

-20

-40
100 101 102 103 104
Frequency (Hz)
Fig. 12 TL curves for ten-layer laminated composite shell with different material

Table 4 Orthotropic materials

Name
Graphite/epoxy Glass/epoxy Boron/epoxy
Density (kg/m3 1,600 1,900 1,600
E 1 (GPa) 137.9 38.6 206
E 2 (GPa) 8.96 8.2 20.6
G 12 (GPa) 7.1 4. 2 6.89
Poisson’s ratio 0.3 0.26 0.3

120
h=0.795 mm
100 h=1.59 mm
h=3.18 mm

80

60
TL (dB)

40

20

-20

-40
0 1 2 3 4
10 10 10 10 10
Frequency (Hz)
Fig. 13 TL curves for ten-layer laminated composite shell with respect to shell thickness

of materials, the TL curves are ascending. Therefore, the TL of glass/epoxy is upper than the others in the
mass-controlled region.
As shown in Fig. 13, changing the thickness of the shell has a broadband effect on TL over the entire range
of the frequency. In general, TL increases about 6 dB as the thickness doubles, which is well anticipated. In a
practical situation, the shells have to be designed as thick as possible, unless there is a weight constraint. The
type of analysis developed in this work will be very useful in such a situation. For example, if the target TL is
known from the consideration of the noise level, the proper thickness of the shell can be easily calculated.
Sound transmission through laminated composite cylindrical shells using analytical model 377

100
R=0.915 m
R=1.83 m
80
R=3.66 m

60

TL (dB)
40

20

-20

-40
0 1 2 3 4
10 10 10 10 10
Frequency (Hz)
Fig. 14 TL curves for ten-layer laminated composite shell with respect to shell radius

100
[0/90/0/90/0]S
[90/0/90/0/90]S
80

60
TL (dB)

40

20

-20

-40
0 1 2 3 4
10 10 10 10 10
Frequency (Hz)
Fig. 15 TL curves for the ten-layered composite shell with respect to stacking sequence

As indicated in Fig. 14, a smaller radius makes the TL higher, which is caused by the curvature effect of
the shell on its stiffness. In mass-controlled region, as a result of shortening the wavelength, all curves show
the same results.
The effect of stacking sequence is shown in Fig. 15. Two patterns [0, 90, 0, 90, 0]s and [90, 0, 90, 0, 90]s
are defined to designate stacking sequence of plies. Since the first pattern contain more plies of 0◦ which place
further away from reference surface leading to higher axial extensional and flexural stiffness, the TL is more
desirable.
The effect of different structural loss factors is shown in Fig. 16. The structural damping becomes unim-
portant in the study of TL except at ring and coincidence frequencies where magnitude of TL increases with
increase of the loss factor near these frequencies.

11 Concluding remarks

In this paper, we studied sound transmission through an infinite laminated composite cylindrical shell. An
analytical expression is derived to calculate the TL including the effect of the external airflow. The frequency
378 K. Daneshjou et al.

100
η= 0.01
80
η= 0.05
η = 0.1

60

TL (dB)
40

20

-20

-40
0 1 2 3 4
10 10 10 10 10
Frequency (Hz)
Fig. 16 TL curves for the ten-layered composite shell with respect to damping

spectrum of TL versus frequency is studied by varying parameter of fluids and shells. The following conclusions
can be drawn from this numerical study:
1. Decreasing of incident angle tends to enhance the TL of cylinder in stiffness-controlled region (below the
ring frequency) and the coincidence frequency is shifted downwards.
2. In higher altitude, acoustic impedance mismatch increases. Therefore, TL in all broadband of frequency is
enhanced.
3. By increasing Mach number, the TL is descending in stiffness-controlled region and the coincidence fre-
quency is shifted upwards.
4. Comparison between aluminum and laminated composite shell is done. It appears that the composite shell
behaves better in noise attenuation below the ring frequency, but it does not appear to be effective as an
aluminum shell in mass-controlled region.
5. The noise attenuation of a composite shell is sensitive to the angle of warp of layers.
6. The structural damping becomes unimportant in the study of TL except at resonance frequencies.

References

1. Beranek, L.L.: Noise and Vibration Control Engineering, Principle and Applications. Wiley, New York (1992)
2. Blaise, A., Lesueur, C., Gotteland, M., Barbe, M.: On sound transmission into an orthotropic infinite shell: comparison with
Koval result and understanding the phenomena. J. Sound Vib. 150, 233–243 (1991)
3. Blaise, A., Lesueur, C.: Acoustic transmission through a 2-D orthotropic multi-layers infinite shell. J. Sound Vib. 155(1),
95–109 (1992)
4. Blaise, A., Lesueur, C.: Acoustic transmission through a 3-D orthotropic multi-layered infinite cylindrical shell, part I:
formulation of the problem. J. Sound Vib. 171(5), 651–664 (1994)
5. Fahy, F.: Sound and Structural Vibration: Radiation, Transmission and Response. Academic, New York (1985)
6. Howe, M.S.: Acoustics of Fluid–Structure Interaction. Cambridge University Press, Cambridge (1998)
7. Gerald, F., Wheatley, P.O.: Applied Numerical Analysis. Addison-Wesley, New York (1997)
8. Kinsler, L.E., Frey, A.R., Coppens, A.B., Sanders, J.V.: Fundamentals of Acoustics, 4th edn. Wiley, New York (2000)
9. Koval, L.R.: On sound transmission into a thin cylindrical shell under flight conditions. J. Sound Vib. 48, 265–275 (1976)
10. Koval, L.R.: On sound transmission into an orthotropic shell. J. Sound Vib. 63, 51–59 (1979)
11. Koval, L.R.: Sound transmission into a laminated composite cylindrical shell. J. Sound Vib. 71(4), 523–530 (1980)
12. Lee, J.H., Kim, J.: Study on sound transmission characteristics of a cylindrical shell using analytical and experimental models.
Appl. Acoust. 64, 611–632 (2003)
13. Leissa, W.A.: Vibration of Shells. NASA, Washington, DC (1973)
14. Mclachlan, N.W.: Bessel Functions for Engineers, 2nd edn. Oxford University Press, Oxford (1955)
15. Norton, M., Karczub, D.: Fundamentals of Noise and Vibration Analysis for Engineers, 2nd edn. Cambridge University
press, Cambridge (2003)
16. Qatu, M.S.: Vibration of Laminated Shells and Plates. Elsevier Academic, Amsterdam (2004)
17. Roussos, L.A., Powell, C.R., Grosveld, F.W., Koval, L.R.: Noise transmission characteristics of advanced composite structure
materials. J. Aircr. 21(7), 528–535 (1984)
Sound transmission through laminated composite cylindrical shells using analytical model 379

18. Reddy, J.N.: Mechanics of Laminated Composite Plates and Shells: Theory and Analysis, 2nd edn. CRC Press, Boca Raton
(2004)
19. Soedel, W.: Vibrations of Shells and Plates. Marcel Dekker, New York (1993)
20. Smith, P.W.: Sound transmission through thin cylindrical shells. J. Acoust. Soc. Am. 29, 721–729 (1957)
21. Tang, Y.Y., Robinson, J.H., Silcox, R.J.: Sound transmission through a cylindrical sandwich shell with honeycomb core. In:
34th AIAA Aerospace Science Meeting and Exhibit (1996)
22. White, P.: Sound transmission through a finite, closed, cylindrical shell. J. Acoust. Soc. Am. 40, 1124–1130 (1966)

S-ar putea să vă placă și