• Aucun résultat trouvé

Buckling and free vibration analysis of laminated composite plates using an efficient and simple higher order shear

N/A
N/A
Protected

Academic year: 2022

Partager "Buckling and free vibration analysis of laminated composite plates using an efficient and simple higher order shear"

Copied!
14
0
0

Texte intégral

(1)

c AFM, EDP Sciences 2016 DOI:10.1051/meca/2015112 www.mechanics-industry.org

M echanics

& I ndustry

Buckling and free vibration analysis of laminated composite plates using an efficient and simple higher order shear

deformation theory

Belkacem Adim

1,4

, Tahar Hassaine Daouadji

1,3,a

, Boussad Abbes

2

and A. Rabahi

1,4

1 D´epartement de g´enie civil, Universit´e Ibn Khaldoun Tiaret, BP 78 Zaaroura, 14000 Tiaret, Alg´erie

2 Laboratoire GRESPI, Campus du Moulin de la Housse, BP 1039, 51687 Reims Cedex 2, France

3 Laboratoire de G´eomatique et D´eveloppement Durable, Universit´e Ibn Khaldoun, Tiaret, Alg´erie

4 Laboratoire de Technologie Industrielle, Universit´e Ibn Khaldoun, Tiaret, Alg´erie Received 13 February 2015, Accepted 21 November 2015

Abstract – In this paper, the buckling and free vibration analysis of laminated composite plates using an efficient and simple higher order shear deformation theory are examined by using a refined shear deformation theory. This theory is based on the assumption that the transverse displacements consist of bending and shear components where the bending components do not contribute to shear forces, and likewise, the shear components do not contribute to bending moments. The most interesting feature of this theory is that it allows for parabolic distributions of transverse shear stresses across the plate thickness and satisfies the conditions of zero shear stresses at the top and bottom surfaces of the plate without using shear correction factors. The number of independent unknowns in the present theory is four, as against five in other shear deformation theories. In this analysis, the equations of motion for simply supported thick laminated rectangular plates are derived and obtained through the use of Hamilton’s principle. The closed- form solutions of anti-symmetric cross-ply and angle- ply laminates are obtained using Navier solution.

Numerical results of the present study are compared with three-dimensional elasticity solutions and results of the first-order and the other higher-order theories reported in the literature. It can be concluded that the proposed theory is accurate and simple in solving the buckling and free vibration behaviors of laminated composite plates.

Key words: Analytical solutions / laminated composite plates / higher-order shear deformation theory / buckling / free vibration / Navier solution

1 Introduction

Laminated composite plates are widely used in indus- try and new fields of technology. Due to the high degrees of anisotropy and the low rigidity in transverse shear of the plates, the Kirchhoff hypothesis as a classical the- ory is no longer adequate. The hypothesis states that the normal to the midplane of a plate remains straight and normal after deformation because of the negligible trans- verse shear effects. Refined theories without this assump- tion have been used recently. The classical laminate plate theory CLPT underpredicts deflections and over predicts frequencies as well as buckling loads with moderately thick plates. Many shear deformation theories account- ing for transverse shear effects have been developed to overcome the deficiencies of the CLPT. The first-order

a Corresponding author:daouadjitah@yahoo.fr

shear deformation theories FSDT based on Reissner [1]

and Mindlin [2] account for the transverse shear effects by the way of linear variation of in-plane displacements through the thickness. A number of shear deformation theories have been proposed to date. The first such the- ory for laminated isotropic plates was apparently [3]. This theory was generalized to laminated anisotropic plates in reference [4]. It was shown in references [5–7], the FSDT violates equilibrium conditions at the top and bot- tom faces of the plate, shear correction factors are re- quired to rectify the unrealistic variation of the shear strain/stress through the thickness. In order to overcome the limitations of FSDT, higher-order shear deformation theories HSDT, since which involve higher-order terms in Taylor’s expansions of the displacements in the thickness coordinate, were developed by Librescu [8], Levinson [9], Bhimaraddi and Stevens [10], Reddy [11], Ren [12], Kant and Pandya [13], and Mohan et al. [14]. A good review

(2)

of these theories for the analysis of laminated composite plates is available in references [15–19]. A refined plate theory using only two unknown functions was developed by Shimpi [20] for isotropic plates, and was extended by Shimpi and Patel [21,22] for orthotropic plates. The most interesting feature of this theory is that it does not re- quire shear correction factors, and has strong similari- ties with the classical plate theory in some aspects such as governing equation, boundary conditions and moment expressions.

In this paper, a refined and simple theory of plates is presented and applied to the investigation of buck- ling and free vibration behavior of laminated composite plates. This theory is based on the assumption that the in-plane and transverse displacements consist of bending and shear components where the bending components do not contribute to shear forces, and likewise, the shear components do not contribute to bending moments. The most interesting feature of this theory is that it allows for parabolic distributions of transverse shear stresses across the plate thickness and satisfies zero shear stress condi- tions at the top and bottom surfaces of the plate without using shear correction factors. The equations of motion are derived using Hamilton’s principle. The fundamental frequencies are found by solving an eigenvalue equation.

The results obtained by the present method are compared with solutions and results of the first-order and the other higher-order theories.

2 Refined plate theory for laminated composite plates

2.1 Basic assumptions

Consider a rectangular plate of total thickness h com- posed of n orthotropic layers with the coordinate system as shown in Figure1. Assumptions of the refined plate’s theory are as follows:

The displacements are small in comparison with the plate thickness and, therefore, strains involved are infinitesimal.

The transverse displacementwincludes three compo- nents of bendingwband shearws. These components are functions of coordinatesx,y, and time tonly.

w(x, y, t) =wb(x, y, t) +ws(x, y, t) (1) The transverse normal stressσz is negligible in com-

parison with in-plane stressesσxandσy.

The displacements U in x-direction and V in y- direction consist of extension, bending, and shear components:

U =u+ub+us, V =v+vb+vs (2) The bending components ub and vb are assumed to be similar to the displacements given by the

classical plate theory. Therefore, the expression for ub andvbcan be given as:

ub=−z∂wb

∂x , vb=−z∂wb

∂y (3a)

The shear componentsusandvsgive rise, in con- junction with ws, to the parabolic variations of shear strains γxz, γyz and hence to shear stresses σxz,σyz through the thickness of the plate in such a way that shear stressesσxz, σyz are zero at the top and bottom faces of the plate. Consequently, the expression forusandvscan be given as:

us=f(z)∂ws

∂x, vs=f(z)∂ws

∂y (3b)

2.2 Kinematics

Based on the assumptions made in the preceding section, the displacement field can be obtained using Equations (1)–(3) as:

u(x, y, z, t) =u(x, y, t)−z∂wb

∂x +f(z)∂ws

∂x v(x, y, z, t) =v(x, y, t)−z∂wb

∂y +f(z)∂ws

∂y

w(x, y, z, t) =wb(x, y, t) +ws(x, y, t) (4a) where u and v are the mid-plane displacements of the plate in the x and y direction, respectively; wb and ws are the bending and shear components of transverse dis- placement, respectively, whilef(z) represents shape func- tions determining the distribution of the transverse shear strains and stresses along the thickness and is given as the present model; the function f(z) is an hyperbolic shape function (Hyperbolic Shear Deformation Theory):

f(z) =z

1 + 3π 2 sech2

1 2

3π 2 htanh

z h

(4b) It should be noted that unlike the first-order shear de- formation theory, this theory does not require shear cor- rection factors. The strains associated with the displace- ments in Equation (4) are:

εx=ε0x+zkxb+fkxs εy=ε0y+zkby+fksy γxy=γxy0 +zkxyb +fkxys γyz=syz

γxz=sxz

εz= 0 (5)

(3)

Fig. 1.Coordinate system and layer numbering used for a typical laminated plate.

where:

ε0x=∂u

∂x, kxb=−∂2wb

∂x2 , kxs =−∂2ws

∂x2 , ε0y =∂v

∂y, kyb=−∂2wb

∂y2 , ksy=−∂2ws

∂y2 γxy0 =∂u

∂y + ∂v

∂x, kxyb =−2∂2wb

∂x∂y, kxys =−2∂2ws

∂x∂y, γyzs = ∂ws

∂y , γxzs = ∂ws

∂x (6) g(z) = 1−f(z) andf(z) =df(z)dz .

2.3 Constitutive equations

Under the assumption that each layer possesses a plane of elastic symmetry parallel to the x-y plane, the constitutive equations for a layer can be written as

⎧⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

⎪⎪

⎪⎪

σx σy σxy σyz σxz

⎫⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎬

⎪⎪

⎪⎪

⎪⎪

⎪⎪

=

⎢⎢

⎢⎢

⎢⎢

⎢⎢

Q11 Q12 0 0 0 Q12 Q22 0 0 0

0 0 Q66 0 0

0 0 0 Q44 0

0 0 0 0 Q55

⎥⎥

⎥⎥

⎥⎥

⎥⎥

⎧⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

⎪⎪

⎪⎪

εx εy γxy γyz γxz

⎫⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎬

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎭ (7)

where Qij are the plane stress-reduced stiffnesses, and are known in terms of the engineering constants in the material axes of the layer:

Q11= E11

1−ν12ν21, Q22= E22

1−ν12ν21, Q12= ν12E22 1−ν12ν21, Q66=G12, Q44=G23, Q55=G13 (8)

Since the laminate is made of several orthotropic layers with their material axes oriented arbitrarily with respect to the laminate coordinates, the constitutive equations of each layer must be transformed to the laminate coordi- nates (x, y, z). The stress-strain relations in the laminate coordinates of the kth layer are given as

⎧⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

⎪⎪

⎪⎪

σx σy σxy σyz σxz

⎫⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎬

⎪⎪

⎪⎪

⎪⎪

⎪⎪

(k)

=

⎢⎢

⎢⎢

⎢⎢

⎢⎢

Q¯11 Q¯12 Q¯16 0 0 Q¯12 Q¯22 Q¯26 0 0 Q¯16 Q¯26 Q¯66 0 0 0 0 0 Q¯44 Q¯45 0 0 0 Q¯45 Q¯55

⎥⎥

⎥⎥

⎥⎥

⎥⎥

(k)

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

εx εy γxy γyz γxz

⎫⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎬

⎪⎪

⎪⎪

⎪⎪

⎪⎪

(k)

(9)

(4)

⎧⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎩

⎧⎪

⎪⎪

⎪⎪

⎪⎩ Nx

Ny

Nxy

⎫⎪

⎪⎪

⎪⎪

⎪⎭

⎧⎪

⎪⎪

⎪⎪

⎪⎩ Mxb Myb Mxyb

⎫⎪

⎪⎪

⎪⎪

⎪⎭

⎧⎪

⎪⎪

⎪⎪

⎪⎩ Mxs Mys

Mxys

⎫⎪

⎪⎪

⎪⎪

⎪⎭

⎫⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎭

=

⎢⎢

⎢⎢

⎢⎢

⎢⎢

⎢⎢

⎢⎢

⎢⎢

⎢⎢

⎢⎢

⎢⎢

⎢⎣

⎢⎢

⎢⎣

A11 A12 A16

A12 A22 A26

A16 A26 A66

⎥⎥

⎥⎦

⎢⎢

⎢⎣

B11 B12 B16

B12 B22 B26

B16 A26 A66

⎥⎥

⎥⎦

⎢⎢

⎢⎣

Bs11 Bs12 Bs16 Bs12 Bs22 Bs26

Bs16 Bs26 Bs66

⎥⎥

⎥⎦

⎢⎢

⎢⎣

B11 B12 B16

B12 B22 B26

B16 A26 A66

⎥⎥

⎥⎦

⎢⎢

⎢⎣

D11 D12 D16

D12 D22 D26

D16 D26 D66

⎥⎥

⎥⎦

⎢⎢

⎢⎣

D11s Ds12 Ds16

D12s Ds22 Ds26 D16s Ds26 Ds66

⎥⎥

⎥⎦

⎢⎢

⎢⎣

Bs11 Bs12 B16s Bs12 Bs22 B26s Bs16 Bs26 B66s

⎥⎥

⎥⎦

⎢⎢

⎢⎣

D11s D12s D16s D12s D22s D26s D16s D26s D66s

⎥⎥

⎥⎦

⎢⎢

⎢⎣

H11s H12s H16s H12s H22s H26s H16s H26s H66s

⎥⎥

⎥⎦

⎥⎥

⎥⎥

⎥⎥

⎥⎥

⎥⎥

⎥⎥

⎥⎥

⎥⎥

⎥⎥

⎥⎥

⎥⎦

⎧⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎩

⎧⎪

⎪⎪

⎪⎪

⎪⎩ ε0x

ε0y γxy0

⎫⎪

⎪⎪

⎪⎪

⎪⎭

⎧⎪

⎪⎪

⎪⎪

⎪⎩ kbx

kby

kxyb

⎫⎪

⎪⎪

⎪⎪

⎪⎭

⎧⎪

⎪⎪

⎪⎪

⎪⎩ kxs

kys kxys

⎫⎪

⎪⎪

⎪⎪

⎪⎭

⎫⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎭

(14a)

⎧⎨

Qsyz Qsxz

⎫⎬

⎭=

As44As45 As45As55

⎧⎨

γyzs γxzs

⎫⎬

⎭ (14b)

where ¯Qijare the transformed material constants given as Q¯11=Q11cos4θ+2(Q12+ 2Q66) sin2θcos2θ+Q22sin4θ Q¯12= (Q11+Q224Q66) sin2θcos2θ+Q12(sin4θ+cos4θ) Q¯22=Q11sin4θ+2(Q12+2Q66) sin2θcos2θ+Q22cos4θ Q¯16= (Q11−Q122Q66) sinθcos3θ

+ (Q12−Q22+ 2Q66) sin3θcosθ Q¯26= (Q11−Q122Q66) sin3θcosθ

+ (Q12−Q22+ 2Q66) sinθcos3θ Q¯66= (Q11+Q222Q122Q66) sin2θcos2θ

+Q66(sin4θ+ cos4θ) Q¯44=Q44cos2θ+Q55sin2θ Q¯45= (Q55−Q44) cosθsinθ

Q¯55=Q55cos2θ+Q44sin2θ (10) In whichθis the angle between the globalx-axis and the localx-axis of each lamina.

2.4 Governing equations

The strain energy of the plate can be written as U =1

2

V

σijεijdV = 1 2

V

xεx+σyεy

+σxyγxy+σyzγyz+σxzγxz)dV (11) Substituting Equations (5) and (9) into Equation (11) and integrating through the thickness of the plate, the strain

energy of the plate can be rewritten as U = 1

2

A

Nxε0x+Nyε0y+Nxyγxy0 +Mxbkbx+Mybkyb +Mxybkxyb +Mxskxs+Myskys+Mxys kxys +Qsyzγyzs

+Qsxzγxzs }dxdy (12)

where the stress resultantsN,M, and Qare defined by (Nx, Ny, Nxy) =

h/2

−h/2

x, σy, σxy)dz

= N k=1

zk+1

zk

x, σy, σxy)dz

(Mxb, Myb, Mxyb) = h/2

−h/2

x, σy, σxy)zdz

= N k=1

zk+1

zk

x, σy, σxy)zdz

(Mxs, Mys, Mxys ) = h/2

−h/2

x, σy, σxy)fdz

= N k=1

zk+1

zk

x, σy, σxy)fdz

(Qsxz, Qsyz) = h/2

−h/2

xz, σyz)gdz

= N k=1

zk+1

zk

xz, σyz)gdz (13)

Substituting Equation (9) into Equation (13) and inte- grating through the thickness of the plate, the stress resultants are given as:

See equations (14a) and (14b) above

(5)

whereAij,Bij, etc., are the plate stiffnesses, defined by Aij, Bij, Dij, Bijs, Dsij, Hijs

= h/2

−h/2

Q¯ij(1, z, z2, f(z), zf(z), f2(z))dz,(i, j) = (1,2,6) (15a) Asij =

h/2

−h/2

Q¯ij[g(z)]2dz,(i, j) = (4,5) (15b)

The work done by applied forces can be written as:

V = 1 2

A

Nx02(wb+ws)

∂x2 +Ny02(wb+ws)

∂y2 +2Nxy0 2(wb+ws)

∂x∂y

dxdy (16) whereNx0,Ny0 andNxy0 are in-plane distributed forces.

The kinetic energy of the plate can be written as T = 1

2

V

ρ¨uiidV=1 2

A

δu

I1u¨−I2∂w¨b

∂x w−I4∂w¨s

∂x

+δv

I1v¨−I2∂w¨b

∂y −I4∂w¨s

∂y

+δwb

I1( ¨wb+ ¨ws) +I2 ∂u¨

∂x +∂v¨

∂y

−I3 2w¨b

∂x2 +2w¨b

∂y2

−I5 2w¨s

∂x2 +2w¨s

∂y2

+δws

I1( ¨wb+ ¨ws) +I4 ∂u¨

∂x +∂v¨

∂y

−I5 2w¨b

∂x2 +2w¨b

∂y2

−I6 2w¨s

∂x2 +2w¨s

∂y2 dxdy (17) whereρis the mass of density of the plate andIiare the inertias defined by

(I1, I2, I3, I4, I5, I6) = h/2

−h/2

ρ(1, z, z2, f(z), zf(z),[f(z)]2)dz (18) Hamilton’s principle [18] is used herein to derive the equa- tions of motion appropriate to the displacement field and the constitutive equation. The principle can be stated in analytical form as

t

0

δ(U +V −T)dt= 0 (19) whereδindicates a variation with respect toxandy.

Substituting Equations (12), (16) and (17) into Equa- tion (19) and integrating the equation by parts, collecting

the coefficients of δu, δv,δwb and δws, the equations of motion for the laminate plate are obtained as follows:

δu: ∂Nx

∂x +∂Nxy

∂y =I1u¨−I2∂w¨b

∂x −I4∂w¨s

∂x δv: ∂Nxy

∂x +∂Ny

∂y =I1v¨−I2∂w¨b

∂y −I4∂w¨s

∂y δwb: 2Mxb

∂x2 + 22Mxyb

∂x∂y +2Myb

∂y2 +N(w)

=I1( ¨wb+ ¨ws) +I2 ∂¨u

∂x+∂¨v

∂y

−I3 2w¨b

∂x2 +2w¨b

∂y2

−I5 2w¨s

∂x2 +2w¨s

∂y2

δws: 2Mxs

∂x2 +22Mxys

∂x∂y +2Mys

∂y2 +∂Qsxz

∂x +∂Qsyz

∂y +N(w)

=I1( ¨wb+ ¨ws) +I4 ∂¨u

∂x+∂¨v

∂y

−I5 2w¨b

∂x2 +2w¨b

∂y2

−I6 2w¨s

∂x2 +2w¨s

∂y2

(20) where N(w) is defined by

N(w) =Nx02(wb+ws)

∂x2 +Ny02(wb+ws)

∂y2

+ 2Nxy0 2(wb+ws)

∂x∂y (21) Equation (20) can be expressed in terms of displacements (u,v,wb,ws) by substituting for the stress resultants from Equation (14). For homogeneous laminates, the equations of motion (20) take the form

See equations (22a)–(22d) next page.

3 Analytical solutions

3.1 Analytical solutions for antisymmetric cross-ply laminates

The Navier solutions can be developed for rectangu- lar laminates with two sets of simply supported boundary conditions. For antisymmetric cross-ply laminates, the following plate stiffnesses are identically zero:

A16=A26=D16=D26=D16s =D26s =H16s =H26s = 0 B12=B26=B16=B66=B12s =B16s =B26s =B66s

=As45= 0

B22=−B11, B22s =−B11s (23)

(6)

A112u

∂x2 + 2A16 2u

∂x∂y+A662u

∂y2 +A162v

∂x2 + (A12+A66) 2v

∂x∂y+A262v

∂y2

−B113wb

∂x3 3B16 3wb

∂x2∂y−(B12+ 2B66) 3wb

∂x∂y2 −B263wb

∂y3

−Bs113ws

∂x3 3B16s 3ws

∂x2∂y−(B12s + 2B66s ) 3ws

∂x∂y2 −B26s 3ws

∂y3 =I1u¨−I2∂w¨b

∂x −I4∂w¨s

∂x (22a)

A162u

∂x2 + (A12+A66) 2u

∂x∂y+A262u

∂y2 +A662v

∂x2 + 2A26 2v

∂x∂y+A222v

∂y2

−B163wb

∂x3 (B12+ 2B66)3wb

∂x2∂y−3B26 3wb

∂x∂y2−B223wb

∂y3

−B16s 3ws

∂x3 (B12s + 2B66s ) 3ws

∂x2∂y 3B26s 3ws

∂x∂y2 −B22s 3ws

∂y3 =I1v¨0−I2∂w¨b

∂y −I4∂w¨s

∂y (22b)

B113u

∂x3 + 3B16 3u

∂x2∂y+ (B12+ 2B66) 3u

∂x∂y2 +B263u

∂y3 +B163v

∂x3 + (B12+ 2B66) 3v

∂x2∂y+ 3B26 3v

∂x∂y2 +B223v

∂y3

−D114wb

∂x4 4D16 4wb

∂x3∂y−2(D12+ 2D66) 4wb

∂x2∂y2 4D26 4wb

∂x∂y3 −D224wb

∂y4

−D11s 4ws

∂x4 4D16s 4ws

∂x3∂y−2(Ds12+ 2D66s ) 4ws

∂x2∂y2 4Ds26 4ws

∂x∂y3 −Ds224ws

∂y4 +N(w)

=I1( ¨wb+ ¨ws) +I2

∂u¨

∂x+∂v¨

∂y

−I3

2w¨b

∂x2 +2w¨b

∂y2

−I5

2w¨s

∂x2 +2w¨s

∂y2

(22c)

B11s 3u

∂x3 + 3B16s 3u

∂x2∂y + (Bs12+ 2B66s ) 3u

∂x∂y2 +B26s 3u

∂y3 +B16s 3v

∂x3 + (B12s + 2B66s ) 3v

∂x2∂y+ 3Bs26 3v

∂x∂y2 +Bs223v

∂y3

−Ds114wb

∂x4 4D16s 4wb

∂x3∂y−2(Ds12+ 2Ds66) 4wb

∂x2∂y2 4D26s 4wb

∂x∂y3 −D22s 4wb

∂y4

−H11s 4ws

∂x4 4H16s 4ws

∂x3∂y 2(H12s + 2H66s ) 4ws

∂x2∂y2 4H26s 4ws

∂x∂y3 −H22s 4ws

∂y4 +As552ws

∂x2 +As442ws

∂y2 + 2As452ws

∂x∂y+N(w)

=I1( ¨wb+ ¨ws) +I4

∂u¨

∂x+∂¨v

∂y

−I5

2w¨b

∂x2 +2w¨b

∂y2

−I6

2w¨s

∂x2 +2w¨s

∂y2

(22d)

(7)

The following boundary conditions for antisymmetric cross-ply laminates can be written as

v(0, y) =wb(0, y) =ws(0, y) = ∂wb

∂y (0, y)

=∂ws

∂y (0, y) = 0

v(a, y) =wb(a, y) =ws(a, y) =∂wb

∂y (a, y)

=∂ws

∂y (a, y) = 0

Nx(0, y) =Mxb(0, y) =Mxs(0, y) =Nx(a, y) =Mxb(a, y)

=Mxs(a, y) = 0

u(x,0) =wb(x,0) =ws(x,0) = ∂wb

∂x (x,0)

=∂ws

∂x (x,0) = 0

u(x, b) =wb(x, b) =ws(x, b) = ∂wb

∂x (x, b)

=∂ws

∂x (x, b) = 0

Ny(x,0) =Myb(x,0) =Mys(x,0) =Ny(x, b) =Myb(x, b)

=Mys(x, b) = 0 (24)

The boundary conditions in Equation (24) are satisfied by the following expansions

u= m=1

n=1

Umneiωtcos(αx) sin(βy)

v= m=1

n=1

Vmneiωtsin(αx) cos(βy)

wb= m=1

n=1

Wbmneiωtsin(αx) sin(βy)

ws= m=1

n=1

Wsmneiωtsin(αx) sin(βy) (25)

where Umn, Vmn, Wbmn and Wsmn unknown parameters must be determined,ω is the eigen frequency associated with (m, n) the eigen-mode, andα=a andβ= b .

Substituting Equations (23) and (25) into Equa- tion (22), the Navier solution of antisymmetric cross-ply

laminates can be determined from equations

⎢⎢

⎢⎢

⎢⎣

s11 s12 s13 s14 s12 s22 s23 s24 s13 s23 s33+k s34+k s14 s24 s34+k s44+k

⎥⎥

⎥⎥

⎥⎦

⎧⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎩ Umn Vmn Wbmn Wsmn

⎫⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎭

+

⎢⎢

⎢⎢

⎢⎣

m11 0 0 0

0 m22 0 0

0 0 m33 m34

0 0 m34 ms44

⎥⎥

⎥⎥

⎥⎦

⎧⎪

⎪⎪

⎪⎪

⎪⎨

⎪⎪

⎪⎪

⎪⎪

U¨mn V¨mn W¨bmn W¨smn

⎫⎪

⎪⎪

⎪⎪

⎪⎬

⎪⎪

⎪⎪

⎪⎪

=

⎧⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎩ 0 0 0 0

⎫⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎪

⎪⎭ (26)

where

s11=A11α2+A66β2, s12=αβ(A12+A66), s13=−B11α3, s14=−Bs11α3

s22=A66α2+A22β2, s23=B11β3, s24=B11s β3 s33=D11α4+ 2(D12+ 2D662β2+D22β4 s34=Ds11α4+ 2(Ds12+ 2Ds662β2+Ds22β4 s44=H11s α4+ 2(H12s + 2H66s2β2+H22s β4

+As55α2+As44β2

m11=m22=I1, m33=I1+I32+β2)

m34=I1+I52+β2), m44=I1+I62+β2), k=Nx0α2+Ny0β2 (27)

3.2 Analytical solutions for antisymmetric angle-ply laminates

For antisymmetric angle-ply laminates, the following plate stiffnesses are identically zero:

A16=A26=D16=D26=Ds16=Ds26=H16s

=H26s = 0

B11=B12=B22=B66=Bs11=Bs12=Bs22

=B66s =As45= 0 (28)

Références

Documents relatifs

A power law representation for a singular stress field is used to fit the obtained stresses in vicinity of the free edge and the parameters are used to assess the CUF elements and

Following the seminal work of Ritz [1], who presented the free vibration solution of a completely free rectangular plate using a series of multiplications of free-free beam

This section presents the thermal buckling results obtained for monolithic and sandwich plates, derived with and without accounting for the effects of nonlinear

An accurate higher-order theory and c0 finite element for free vibration analysis of laminated composite and

Abstract : In this paper a 2D quadrilateral finite element has been developed based on Reddy’s third order shear deformation theory for the natural vibration behavior analysis

[16,17] developed a four node finite element C1 with five degrees of freedom in each node based on Reddy’s third-order shear deformation theory for the bending

This histogram (Figu 6) represents the rupture factor by deformation and the critical buckling factor of the individual optimal after the generation of the final

(2013) developed an isogeometric fi nite element method for static, free vibration and buckling analysis of laminated composite and sandwich plates based upon layer-wise theory..