You are on page 1of 16

Applied Mathematical Modelling 35 (2011) 5059–5074

Contents lists available at ScienceDirect

Applied Mathematical Modelling


journal homepage: www.elsevier.com/locate/apm

Buckling analysis of shear deformable plates using the quadrature


element method
Hongzhi Zhong ⇑, Chunlin Pan, Hao Yu
Department of Civil Engineering, Tsinghua University, Beijing, 100084, China

a r t i c l e i n f o a b s t r a c t

Article history: The recently proposed weak form quadrature element method (QEM) is applied to buckling
Received 18 July 2010 analysis of shear deformable plates. Integrals involved in the variational description a plate
Received in revised form 12 April 2011 are evaluated first by an efficient numerical integration scheme wherein the partial deriv-
Accepted 15 April 2011
atives at the integration points are approximated using differential quadrature analogs. In a
Available online 24 April 2011
quadrature element, neither the node pattern nor the number of nodes is fixed, being
adjustable according to convergence need. Three examples are presented and comparison
Keywords:
with finite element results is made to demonstrate the effectiveness and computational
Buckling
Differential quadrature
efficiency of the QEM for buckling analysis of shear deformable plates.
Quadrature element method Ó 2011 Elsevier Inc. All rights reserved.
Shear deformable plate
Weak form

1. Introduction

Buckling failure of plates is of concerns in various industrial areas. The typical classical plate theory, also known as the
Kirchhoff plate theory, is restricted to thin plates and becomes inappropriate when the thickness-to-dimensions of plates
becomes appreciable. It is well-known that the utilization of the classical plate theory may overestimate the critical load
of a moderately thick plate. With the increase of plate thickness, the transverse shear deformation of the plate begins to
reduce the overall stiffness of the plate and lower the buckling resistance accordingly. The first-order shear deformable plate
theory, also known as the Reissner–Mindlin plate theory [1,2], has been recognized as a simple way to take into account the
transverse shear effect of moderately thick plates. In practice, the use of numerical tools such as the finite element method is
often the only way to proceed to obtain the critical load and the buckling mode of a structural member for design. The finite
element method has been developed to such an extent that most practical problems can be dealt with satisfactorily. How-
ever, the finite element method uses typically low-order approximations and high computational cost is often needed to
achieve high accuracy of results. Engineers and analysts favor an efficient and simple numerical tool requiring less compu-
tational cost and background knowledge. Therefore, the development of new numerical methods has been an area of sustain-
ing interest to researchers.
Recently, a variation-based method, termed as the weak form quadrature element method (QEM), has been applied to
analysis of various structural problems [3–8]. It starts from evaluation of the integrals involved in the variational description
of a problem with an efficient numerical integration scheme, being followed by approximation with the differential quad-
rature analog of the derivatives at the integration nodes. As a result, high computational efficiency is achieved and less data

⇑ Corresponding author.
E-mail address: hzz@tsinghua.edu.cn (H. Zhong).

0307-904X/$ - see front matter Ó 2011 Elsevier Inc. All rights reserved.
doi:10.1016/j.apm.2011.04.030
5060 H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074

preparation is needed, for the integration points are adjustable to meet convergence requirement. The prefixal phrase, weak
form, is used for distinction from the strong form quadrature element method proposed by Striz et al. [9] which was applied
directly to the solution of differential equations. In the full text of this paper, the term quadrature element refers exclusively
to the weak form approach. Actually, Striz et al. [10] took the initiative to incorporate the differential quadrature analog into
the weak form description of a plate problem. Differing from the present work substantially, however, their node position
and number of nodes are pre-assigned. As such, their approach bears a resemblance to the finite element method and results
in limited enhancement of efficiency and geometric feasibility.
In present investigation, buckling analysis of shear deformable plates is performed using the weak form quadrature ele-
ment method. Convergence rate and CPU time needed in the quadrature element analysis are examined. Three examples are
given to verify the method for buckling analysis of moderately thick plates. The first example is a simply supported rectan-
gular plate subjected to uniform uni-axial compression. A simply supported rectangular plate under uniform shear is studied
in the second example. In the third example, buckling of a square plate with a central circular cut-out is investigated. Results
for both very thin plates and moderately thick plates are compared with the p-type finite element [11] solution and those of
the computer code ANSYS.

2. Theoretical background

2.1. Geometric transformation

Similar to other numerical methods, discretization of the problem domain is the first step in quadrature element analysis
unless the domain of interest is regular and quadrilateral. Usually an arbitrarily-shaped plate is divided into a number of
large quadrilaterals that are feasible for transformation onto the standard computational domain – a bi-unit square. For
an arbitrary quadrilateral domain, a quadrature element, the following Cartesian-standard geometric transformation is
introduced
 P
x ¼ xðn; gÞ ¼ Si ðn; gÞxi ;
P  1 6 n; g 6 1; ð1Þ
y ¼ yðn; gÞ ¼ Si ðn; gÞyi ;

where x and y are space variables defined on the physical domain of the element, n and g are variables defined on the bi-unit
square domain. The transformation functions Si(n, g) are often constructed as the shape functions for serendipity-type finite
elements wherein only boundary nodes are needed for the geometric transformation. Alternatively, the blending function
strategy [12], which is widely used in p-type finite elements, is a choice as long as analytical expressions of curved sides
of a quadrilateral are available.
The discretization of a problem domain and the geometric transformation for a quadrature element will be exemplified by
examining a square plate with a central circular cut-out.
As shown in Fig. 1, a quarter of the plate is usually chosen as the domain to be modeled if the double symmetry of the
domain is exploitable. Suppose that the first quadrant of the plate is considered in quadrature element analysis. Discretiza-
tion is done simply by dividing the first quadrant into two quadrilateral elements with the line bisecting the domain. In a
serendipity-type transformation, a number of side nodes are needed to represent a curved side accurately. The quintic ser-
endipity-type transformation [4], for instance, is a good compromise between accuracy and mathematical convenience. The
quintic serendipity-type transformation for element one in Fig. 1 is given as follows

3 η
4(-1, 1) 3(1, 1)

2 8
7
6 ξ
5
L a o 4 x
1 (0, 0)

1(-1,-1) 2(1,-1)

Fig. 1. A square plate with a central circular cut-out.


H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074 5061

8
1
>
> S5 ðn; gÞ ¼  0:49152 ð1 þ nÞð1  nÞð1  gÞðn þ 0:2Þðn  0:2Þðn  0:6Þ;
>
>
>
< S6 ðn; gÞ ¼ 1 ð1 þ nÞð1  nÞð1  gÞðn þ 0:6Þðn  0:2Þðn  0:6Þ;
0:24576
>
> 1
S7 ðn; gÞ ¼  0:24576 ð1 þ nÞð1  nÞð1  gÞðn þ 0:6Þðn þ 0:2Þðn  0:6Þ;
>
>
>
: 1
S8 ðn; gÞ ¼ 0:49152 ð1 þ nÞð1  nÞð1  gÞðn þ 0:6Þðn þ 0:2Þðn  0:2Þ
8 ð2Þ
> S1 ðn; gÞ ¼ 0:25ð1  nÞð1  gÞ  0:8S5 ðn; gÞ  0:6S6 ðn; gÞ  0:4S7 ðn; gÞ  0:2S8 ðn; gÞ;
>
>
>
< S2 ðn; gÞ ¼ 0:25ð1 þ nÞð1  gÞ  0:2S5 ðn; gÞ  0:4S6 ðn; gÞ  0:6S7 ðn; gÞ  0:8S8 ðn; gÞ;
>
> S3 ðn; gÞ ¼ 0:25ð1 þ nÞð1 þ gÞ;
>
>
:
S4 ðn; gÞ ¼ 0:25ð1  nÞð1 þ gÞ:
As element one is a quadrilateral with three straight sides and one circular arc side, the blending function transformation is
feasible. Note that the subtending angle of the arc is p/4. The blending function transformation can be formulated as
(  
xðn; gÞ ¼ 12 g a cos p8 ð1 þ nÞ þ x3 ð1þnÞ4ð1þgÞ þ x4 ð1nÞ4ð1þgÞ ;
  ð3Þ
yðn; gÞ ¼ 12 g a sin p8 ð1 þ nÞ þ y3 ð1þnÞ4ð1þgÞ þ y4 ð1nÞð1þ
4

:
Obviously, the blending function transformation is more convenient and accurate than the serendipity-type point transfor-
mation. Similar geometric transformation can be made for element two in Fig. 1.
From Eq. (1), the Jacobian matrix and its inverse are obtained as follows
" # " #  
xn yn 1
nx gx 1 yg yn
J¼ ; J ¼ ¼ ; ð4Þ
xg yg ny gy jJj xg xn

where the Jacobian determinant is


jJj ¼ xn yg  xg yn : ð5Þ

Then, the transformation for the first order partial derivative is expressed as
( ) ( @
)
@
@x 1 @n
@ ¼J @
: ð6Þ
@y @g

which will be used during the evaluation of integrals in the weak form description of first order shear deformable plates.

2.2. Weak form description of moderately thick plates

The material of the plate is assumed to be homogeneous, isotropic, and linearly elastic, and the applied forces to be con-
servative. The displacement field of the plate (see Fig. 2) is assumed to take the following form
8
< ux ðx; y; zÞ ¼ uðx; yÞ  ux ðx; yÞz;
>
uy ðx; y; zÞ ¼ v ðx; yÞ  uy ðx; yÞz; ð7Þ
>
:
uz ðx; y; zÞ ¼ wðx; yÞ;
where u(x, y), v(x, y), w(x, y) are the three translational displacement components along the three Cartesian coordinate axes at
an arbitrary point on the mid-surface, and ux(x, y), uy(x, y) are angular rotations about y and x axes, respectively. The elastic
material law for the plate is written as
r ¼ Ee ð8Þ

w
ϕy x, ux

ϕx
y, uy
h

z, uz

Fig. 2. Displacement components of a shear deformable plate.


5062 H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074

with

rT ¼ ½ rx ry sxy sxz syz ;


: ð9Þ
eT ¼ ½ ex ey exy exz eyz 

E is a symmetric square matrix of order five, involving Young’s modulus E and Poisson’s ratio l of the material. The non-zero
entries of E are given as

E lE E
E11 ¼ E22 ¼ ; E12 ¼ E21 ¼ ; E33 ¼ E44 ¼ E55 ¼ : ð10Þ
1  l2 1  l2 2ð1 þ lÞ

The strains of the first order shear deformation plate can be written as
e ¼ e þ e; ð11Þ

where e is the linear Green strain components, and e is the nonlinear components. The former is given by
8 @u ux 9
>
> @x
 z @@x >
>
>
> >
>
>
> @v @ uy >
>
>
> @y  z >
>
h iT < 
@y
   =
@ux @uy @ux @uy @uz @uy @ uy
e¼ þ þ @u x @uz
þ @z ¼ @u
þ @ v  z @ ux
þ : ð12Þ
@x @y @y @x @x @z @y >
> @y @x @y @x >
>
>
> >
>
>
> @w
 ux >
>
>
> @x >
>
: @w ;
@y
 u y

The latter is partitioned into two portions: one associated with translational displacements and the other associated with
rotational displacements, i.e.

e ¼ e1 þ e2 ; ð13Þ

where
8 h 2  2  2 i 9
>
>
1 @u
þ @@xv þ @w >
>
> 2 @x
> @x >
>
>
>  2  2  2 >  >
>
> 1 @u >
>
> @v
< 2 @y þ @y þ @y @w >
=
e1 ¼ ð14Þ
>
> @u @u
þ @@xv @@yv þ @w @w >
>
>
> @x @y @x @y >
>
>
>0 >
>
>
> >
>
>
: >
;
0
and
8    2  9
2
>
> 1 @ ux @ uy >
>
>
>2 þ >
>
>
>
@x @x >
>
>
>   >
>  2 @ u 2 >
> >
>
>
<21 @ ux
þ y >
=
@y @y
e2 ¼ z2 : ð15Þ
>
> @ ux @ ux @ uy @ uy >
>
>
> þ >
>
>
> @x @y @x @y >
>
>
> >
>
>
> 0 >
>
>
: >
;
0
It is noteworthy that Eq. (14) is the predominant nonlinear strain, while Eq. (15) is often neglected in most finite element
formulations. The significance to include Eq. (15) will be highlighted later in the second example problem. Besides, the mixed
nonlinear strain terms that involve both translational displacements and the rotational displacements are extraneous and
therefore omitted in Eq. (13) since they vanish after integration over the thickness coordinate z.
The total potential energy of the element is comprised of two parts, i.e.

PðeÞ ¼ U ðeÞ þ V ðeÞ ; ð16Þ


(e) (e)
where U and V are the strain energy and potential for external loads, respectively. The exact change of the total potential
energy is then given in terms of the Taylor expansion as
1 2 ðeÞ
DPðeÞ ¼ dPðeÞ þ d P þ ðnegligible termsÞ: ð17Þ
2!
H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074 5063

For an equilibrium state, the first variation of the total potential energy is stationary, i.e.

dPðeÞ ¼ dU ðeÞ þ dV ðeÞ ¼ 0: ð18Þ


For moderately thick plates that are characterized by small prebuckling displacements, it is reasonable to perform linear
analysis at this stage [13], neglecting nonlinear effects. A fundamental state of the element is then obtained which will
be used to evaluate the buckling load. Suppose that the initial stress state in the element is found to be

r0 ¼ r0x r0y s0xy 0 0 : ð19Þ

The internal forces resulting from the initial stresses in the element are given as
R h=2 R h=2 R h=2
N0x ¼ h=2
r0x dz; N0y ¼ h=2
r0y dz; N0xy ¼ h=2
s0xy dz;
R h=2 R h=2 R h=2 ð20Þ
P0x ¼ h=2
r0x z2 dz; P0y ¼ h=2
r0y z2 dz; P0xy ¼ h=2
s0xy z2 dz
which can be simplified into

N0x ¼ r0x h; N0y ¼ r0y h; N0xy ¼ s0xy h;


3 3 3
ð21Þ
P0x ¼ r0x 12
h
; P0y ¼ r0y 12
h
; P0xy ¼ s0xy 12
h

if a constant stress state prevails in the element.


The internal forces given in (20) or (21) are often regarded as an appropriate approximation for the actual fundamental
state. For simplicity, the second variation of the linear strains and all terms higher than the second order are neglected in
linearized buckling analysis. The buckling criterion is acquired from the observation of the second variation of the total po-
tential energy [14], i.e.

1 2 ðeÞ 1
d P ¼ d2 U ðeÞ ðnote d2 V ðeÞ ¼ 0Þ: ð22Þ
2! 2!
The strain energy is partitioned into two parts, i.e.

U ðeÞ ¼ U ðeÞ þ U ðeÞ ; ð23Þ

where U(e) stands for the strain energy associated with e, the linear strains, while U(e) is the strain energy associated with the
nonlinear terms e.

2.3. Quadrature element formulation

Upon accomplishment of the integration over the z coordinate, the strain energy associated with linear strains can be ex-
pressed as

Z 1 Z 1
1 T T
U ðeÞ ¼ IðeÞ dndg; IðeÞ ¼ v K DKvjJj; ð24Þ
1 1 2
where the auxiliary strain vector is defined as

h iT
@ ux @ ux @ uy @ uy @v @v
v ¼ ux uy @n @g @n @g
@u
@n
@u
@g @n @g
@w
@n
@w
@g
: ð25Þ

From Eq. (6), the transformation matrix for linear strains is given by
2 3
0 0 0 0 0 0 yg yn 0 0 0 0
6 7
60 0 0 0 0 0 0 0 xg xn 0 0 7
6 7
6 7
60 0 0 0 0 0 xg xn yg yn 0 0 7
6 7
6 7
16 0 0 yg yn 0 0 0 0 0 0 0 0 7
K¼ 6 6
7: ð26Þ
jJj 6 0 0 0 0 xg xn 0 0 0 0 0 0 7 7
6 7
60 0 xg xn yg yn 0 0 0 0 0 0 7
6 7
6 7
6 jJj 0 0 0 0 0 0 0 0 0 yg yn 7
4 5
0 jJj 0 0 0 0 0 0 0 0 xg xn
5064 H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074

D is a sparse symmetric matrix of order eight whose non-zero entries are given as

l lÞ
D11 ¼ D22 ¼ 12
h2
D; D12 ¼ D21 ¼ 12
h2
D; D33 ¼ 6ð1
h2
D;
D44 ¼ D55 ¼ D; D45 ¼ D54 ¼ lD; ð27Þ
1l 6ð1lÞj
D66 ¼ 2
D; D77 ¼ D88 ¼ h2
D;

where
3
Eh
D¼ ð28Þ
12ð1  l2 Þ

is the flexural rigidity of the plate and j is the shear correction factor that is taken as 5/6 in the present investigation.
In weak form quadrature analysis, the Lobatto integration scheme is usually chosen to evaluate the integrals involved in
the weak form description of the problem. For convenience, the same number of integration points N is used in the two
directions of the element in the standard square domain. Denoting Wi as the weighting coefficients for the numerical inte-
gration scheme, Eq. (24) is rewritten as
X
N X
N
ðeÞ 1X N X N
U ðeÞ ¼ W i W j Iij ¼ W i W j vTij KTij Dij Kij vij jJjij : ð29Þ
i¼1 j¼1
2 i¼1 j¼1

The Lobatto integration point distribution is

n1 ¼ 1; . . . ; nk ; . . . ; nN ¼ 1; k ¼ 2; . . . ; N  1; ð30Þ

where nk are the (k  1)th zeros of first order derivative of (N  1)th order Legendre polynomial, i.e.

dPN1 ðnÞ
¼ 0: ð31Þ
dn

The weighting coefficients Wi are given as [15]


( 2 2
W 1 ¼ NðN1 Þ
; W N ¼ NðN1 Þ
;
ð32Þ
W k ¼ NðN1Þ½P2 2 ðk ¼ 2; 3; . . . ; N  1Þ:
N1 ðnk Þ

The partial derivatives in Eq. (25) are approximated using the differential quadrature analog which was first proposed by
Bellman and Casti [16]. The essence of the differential quadrature analog lies in that the (partial) derivative of a function
at a point is approximated with weighted linear summation of function values at all discretized points, i.e.

XN
df ðnÞ ð1Þ
¼ C ij f ðnj Þ; i ¼ 1; 2; . . . ; N ð33Þ
dn n¼ni j¼1

ð1Þ
where C ij , weighting coefficients for the first order derivatives, are given by
8
> H0 ðni Þ
< ðni nj ÞH0 ðnj Þ ; ði – jÞ;
>
ð1Þ
C ij ¼ P N ð34Þ
>
> ð1Þ
C ik ; ði ¼ jÞ
:
k¼1;k–i

with
Y
N Y
N
HðnÞ ¼ ðn  nj Þ; H0 ðni Þ ¼ ðni  nj Þ: ð35Þ
j¼1 j¼1;j–i

Details of the differential quadrature analog can be referred to references [17,18]. Denote the nodal displacement vector of a
quadrature element for the plate as
h iT
ðeÞ
d ¼ ðux Þ11 ðuy Þ11 u11 v 11 w11 ðux Þ12 ðuy Þ12 u12 v 12 w12 . . . ðux ÞNN ðuy ÞNN uNN v NN wNN : ð36Þ
H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074 5065

then the vectors vij in Eq. (25) can be expressed as


8 9
>
> u >
>
> xij
> >
>
>
> uyij >
>
>
> >
>
>
> >
>
>
> P ð1Þ
N >
>
>
> C ðu Þ >
x kj >
>
> ik >
>
8 9 > > k¼1 >
>
>
> >
>
>
> ux jij > > >
> PN >
>
>
> >
> >
> C jk ðux Þik >
ð1Þ
>
>
> u j >
> >
> >
>
>
> y ij >
> >
> k¼1 >
>
>
> >
> >
> >
>
>
> @ ux
jij >> >
> PN >
>
>
> @n >
> >
> C ik ðuy Þkj >
ð1Þ
>
>
> >
> >
> >
>
>
> @ u x
j >
> >
> k¼1 >
>
>
> @ g ij >
> >
> >
>
>
> >
> >
> PN >
>
>
> @ uy >
> >
> C
ð1Þ
u >
>
> j ij >
> >
> jk ð y ik >
Þ >
>
>
>
@n >
> >
> k¼1 >
>
>
< @ uy j > = > <P >
=
N
@ g ij ð1Þ ðeÞ
vij ¼ @u ¼ C ik ukj ¼ Bij d ; ð37Þ
>
> j >
> >
> >
>
>
> @n ij > > >
>
k¼1 >
>
>
> > >
> > PN >
>
>
> @u
jij > > >
> ð1Þ >
>
>
> @ g >
> >
> C u ik >
>
> @v >
> > > > k¼1 jk >
>
>
> j >
> >
> >
>
>
> @n ij >
> > >
>P N >
>
>
> >
> > ð1Þ >
>
>
>
@v
j >
> >
> C ik v kj >
>
>
> @ g ij > >
> > k¼1 >
>
>
> >
> >
> >
>
> @w
j
> @n ij > >
> > >
>P N >
>
>
> > > ð1Þ >
>
>
: @w j > ; > >
> C jk v ik >
>
>
@g ij >
> k¼1 >
>
>
> >
>
>P
> N >
>
>
> C
ð1Þ
w >
>
>
> ik kj >
>
> k¼1
> >
>
>
> >
>
>
> P ð1Þ
N >
>
>
>
: C jk wik > >
;
k¼1

where Bij are matrices of size 12  5N2, involving the weighting coefficients for the first order derivatives from (33).
Substitution of Eq. (37) into Eq. (29) yields

1X N X N
ðeÞT ðeÞ 1 ðeÞT ðeÞ
U ðeÞ ¼ W i W j d BTij KTij Dij Kij Bij jJjij d ¼ d KðeÞ d ; ð38Þ
2 i¼1 j¼1 2

where
X
N X
N
KðeÞ ¼ W i W j BTij KTij Dij Kij Bij jJjij : ð39Þ
i¼1 j¼1

In a similar manner, the strain energy associated with the nonlinear strains can be expressed as:
Z 1 Z 1
1 T T
U ðeÞ ¼ IðeÞ dndg; IðeÞ ¼ v K DKvjJj; ð40Þ
1 1 2
where
h iT
@v @v @ ux @ ux @ uy @ uy
v¼ @u
@n
@u
@g @n @g @n @g @n @g
@w
@n
@w
@g
; ð41Þ

2 3
J1 0 0 0 0
6 7
60 J1 0 0 0 7
6 7
K¼6
60 0 J1 0 0 7
7 ð42Þ
6 7
40 0 0 J1 0 5
0 0 0 0 J1
and
2 3
N0 0 0 0 0
6 7
60 N0 0 0 0 7
6 7
D¼6
60 0 P0 0 0 7
7 ð43Þ
6 7
40 0 0 P0 0 5
0 0 0 0 N0
5066 H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074

with
" # " #
N0x N0xy P0x P0xy
N0 ¼ ; P0 ¼ : ð44Þ
N0xy N0y P0xy P0y
Using the same numerical integration scheme as in (29), Eq. (40) turns into
1X N X N
ðeÞT ðeÞ 1 ðeÞT ðeÞ ðeÞ
U ðeÞ ¼ W i W j d BTij KTij Dij Kij Bij jJjij d ¼ d KG d ; ð45Þ
2 i¼1 j¼1 2

where

ðeÞ
X
N X
N
KG ¼ W i W j BTij KTij Dij Kij Bij jJjij : ð46Þ
i¼1 j¼1

In Eq. (45), the differential quadrature analogs of the partial derivatives in Eq. (41) at an integration point have been intro-
duced, i.e.
8 9
>
> PN
ð1Þ >
>
>
> C u kj >
>
>
> ik >
>
>
> k¼1 >
>
>
> >
>
>
> P ð1Þ
N >
>
>
> C u ik >
>
>
> jk >
>
>
> k¼1 >
>
8 @u 9 >
> >
>
> jij > >
> PN
ð1Þ >
>
>
> @n >
> >
> C ik v kj >
>
>
> > > >
>
>
@u
g jij > >
>
>
>
> k¼1 >
>
>
>
>
@ >
> >
> >
>
>
> @v >
> >
> PN >
>
>
> j ij > > >
> C
ð1Þ
v >
>
>
>
@n >
> >
> jk ik >
>
>
> @v
j >
> >
> k¼1 >
>
>
> @g > ij > >
> >
>
>
> > > PN >
>
> ux > > >
> C
ð1Þ
ðu Þ
>
>
< @@n jij >
= >
< ik x kj >
=
k¼1 ðeÞ
vij ¼ @ ux ¼ ¼ Bij d ; ð47Þ
>
> j >
@ g ij >
>
> PN >
>
>
> >
> >
> C jk ðux Þik >
ð1Þ
>
>
> yj >
@ u > >
> k¼1 >
>
>
> > > >
> @n ij > > >
> >
>
>
> @u > > >
> P ð1Þ
N >
>
>
> > > >
>
>
y
j >
> >
> C ðu Þ >
>
> @g ij >
> >
>
>
> k¼1
>
ik y kj >
>
>
>
> @w >
> >
> >
>
>
> @n jij > > >
>P N >
>
>
> > > >
: @w > ; >
>
> C jk ðuy Þik >
ð1Þ
>
>
j
@ g ij > k¼1
> >
>
>
> >
>
>
>P N >
>
>
> >
C ik wkj >
ð1Þ
>
> >
>
>
> k¼1 >
>
>
> >
>
>
> >
>
>
> PN >
C jk wik >
ð1Þ
>
: >
;
k¼1

where Bij are matrices of size 10  5N2, involving the weighting coefficients for first order derivatives from Eq. (33). From
Eqs. (23), (38) and (45), the second variation of the total potential energy is rewritten as
1 2 ðeÞ 1 1 ðeÞT  ðeÞ ðeÞ

ðeÞ
d P ¼ d2 U ðeÞ ¼ dd K þ KG dd ð48Þ
2! 2! 2!
giving the stability criterion
 
ðeÞ
KðeÞ þ KG dd ^ ðeÞ ¼ 0; ð49Þ

where d^ ðeÞ stands for the incremental displacement vector from the fundamental state. Denote U as the total potential energy
of the entire plate which is the summation of all element contributions, i.e.
X
U¼ U ðeÞ : ð50Þ
e

Applying the stability criterion to the entire plate yields the global equations of the system
^ ¼ 0;
ðK þ KG Þdd ð51Þ
where d^ is the global incremental nodal displacement vector; K and K are the global stiffness matrix and the global geo-
G
metric stiffness matrix, respectively. The assemblage of the global matrices K and KG follows the routine finite element pro-
cedure. In linearized buckling analysis, the internal forces are assumed to be proportional to the applied load. A scalar is
often introduced in Eq. (49), i.e.
H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074 5067

^ ¼ 0;
ðK þ kKG0 Þdd ð52Þ
where KG0 is often computed for a unit applied load. Among the eigenvalues of Eq. (52), the minimum one, kcr, is often
sought, which determines the critical load of the plate.

3. Examples and results

Three examples are studied in this section. The first two examples are benchmark problems where theoretical solutions
are available for thin plates [19]. The third example aims at verification of the weak form quadrature element solution for
complicated problems. In all computational cases, the Poisson’s ratio l is taken as 0.30.

3.1. Example one

To illustrate the effectiveness of the quadrature element method, a simply supported rectangular plate under uniform
uniaxial in-plane compression is studied first (see Fig. 3). All sides of the plate are simply supported. Therefore, the essential
boundary conditions at the four sides require that

w ¼ ux ¼ 0; at x ¼ 0; a ð53Þ
and

w ¼ uy ¼ 0; at y ¼ 0; b ð54Þ

To explore the capability of the quadrature element formulated for moderately thick plates, the entire rectangular plate is
modeled by one element despite the double symmetry of the plate. The plate is also modeled by one p-type finite element
for comparison. The commonly adopted non-dimensional critical load coefficient is introduced, i.e.
2
ðNx Þcr b
k1 ¼ ; ð55Þ
p2 D
where (Nx)cr is the critical buckling load of the plate. The theoretical value of k1 for a simply supported thin rectangular plate
under uniform compression is given in [19]. The buckling load for a square plate is computed and compared with those ob-
tained using one p-type plate element as well as the finite element computer code ANSYS with element – Shell181, which
has four nodes at the four corners of the quadrilateral element and five degrees of freedom per node: three translations and
two rotations [20]. The relative errors of the coefficient k1 are shown in Fig. 4. The total number of degrees-of-freedom for a
p-type plate element is calculated by

5ðp þ 1Þðp þ 2Þ
Np ¼ þ 10: ð56Þ
2
It is seen that the QEM solution converges at least as rapidly as that of the p-type FEM and far more rapidly than that of
ANSYS with the increase of the number of degrees-of-freedom. The exact solution is reproduced when the order of approx-
imation N in quadrature element is increased to five. It is noted that the p-type finite element has limitation on its polyno-
mial order p (=N  1), depending on the aspect ratio of the plate. In the case of square plate, the result of p-type finite
element goes wild when p is increased to 13, contrasting with the invariably stable quadrature element solution. The break-
down of higher order p-type finite element solution is due to its seriously ill-conditioned stiffness matrix as enunciated in
detail in [21]. In Fig. 5, the CPU time consumption for the quadrature element solution is compared with that of the p-FEM.
All routines of the present investigation are written in FORTRAN 95 and run on a desktop with an Intel dual core 3.0 GHz
processor and 3.25 GB memory.

x
y

Nx Nx b

a
Fig. 3. A simply supported rectangular plate compressed uniformly in one direction.
5068 H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074

Fig. 4. Relative error of buckling coefficient of a simply supported square plate under uniformly uniaxial compression.

Fig. 5. CPU time for numerical solutions of a simply supported square plate compressed uniformly.

It is found that the QEM enjoys high efficiency in comparison with the p-type finite element, consuming roughly half of
the CPU time for the p-type finite element analysis. This is ascribed to the unified node set in the quadrature element anal-
ysis while high-order Gauss quadrature in the p-type finite element solution process is rather time-consuming. Admittedly,
the total consumption of CPU time for a problem of this size is minute regardless of either numerical method but it may be of
concern when a large-scale problem or structural optimization is considered. More computed results of buckling load for
rectangular plates with a variety of aspect ratios are given in Fig. 6. Being consistent with the results in the literature, the
buckling coefficient gets lower with the increase of thickness-to-width ratio.
In addition, it is noted that the p-type finite element solution of the problem breaks down when the aspect ratio a/b is
larger than 3.4 for thin plates regardless of the increase of approximation order in the element. The convergent criterion
for both the quadrature element solution and the p-type finite element solution is set as

ðNÞ ðN1Þ
ðk1 Þ  ðk1 Þ
< 0:01%; ð57Þ
ðk1 ÞðNÞ
H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074 5069

Fig. 6. Buckling coefficient of a simply supported rectangular plate versus plate aspect ratio.

where the numerator represents the solution difference for two adjacent approximation orders. The non-convergent situa-
tion for the p-type finite element solution becomes worse off with the increase of the plate thickness. For instance, the larg-
est aspect ratio of the plate modeled successfully using one p-type element is reduced to 3.2 when thickness-to-width ratio
is increased to 0.1. As aforementioned, it is the ill-conditioned stiffness matrix of the p-type finite element [21] that fails the
solution. The problem can be overcome if the plate is divided into more p-type elements. In contrast, one quadrature element
solution suffices to meet the criterion in Eq. (57) regardless of the thickness-to-width ratio and the plate aspect ratio.

3.2. Example two

In this example, buckling analysis is conducted for a simply supported rectangular plate under uniform shear load along
the four sides (see Fig. 7). The same material and geometric parameters as in example one are used. A dimensionless eigen-
buckling load coefficient is defined as
2
ðNxy Þcr b
k2 ¼ : ð58Þ
p2 D
The theoretical solution for a simply supported thin rectangular plate under shear is given in [19]. The variation of k2 with
the change of number of degrees-of-freedom is displayed in Fig. 8 for both one quadrature element solution and one p-type
finite element solution. Convergence of the quadrature element solution is attained when N is increased to seven. Again, the

x Nyx
y

Nxy b

a
Fig. 7. A simply supported rectangular plate under uniform shear.
5070 H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074

Fig. 8. Relative errors of numerical solutions and the buckling mode of a simply supported square plate under shear.

p-type finite element solution exhibits divergence when the order of the approximation is over 13. The significance of Eq.
(15) is examined by checking the relative error of k2 resulting from omission of the nonlinear strains in Eq. (15). From
Fig. 9, it is found that the relative error of k2 increases with the thickness-to-width ratio. For a square plate, the relative error
is larger than 5% when the thickness-to-width ratio is large than 0.12. Clearly, the nonlinear strain terms in Eq. (15) should be
retained in buckling analysis of plates especially for thicker plates. Fig. 10 shows the CPU time consumption, highlighting the
high efficiency of the QEM in comparison with that of the p-type finite element solution. More results for the shear buckling
coefficient k2 for rectangular plate with various aspect ratios are shown in Fig. 11. It is seen that the exact solution is virtually
reproduced by the QEM and the p-type finite element predictions when the aspect ratio of the plate is small. It is found that
the one p-type finite element solution breaks down when the aspect ratio of the plate is larger than 1.8 for moderately thick
plate (h/b = 0.1) and larger than 2.0 for thin plates (h/b = 0.001). Again, the buckling coefficient gets lower with the increase
of thickness-to-width ratio.

Fig. 9. Effect of neglecting nonlinear strain terms associated with rotational displacements in Eq. (15).
H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074 5071

Fig. 10. CPU time for numerical solutions of a simply supported square plate sheared uniformly.

Fig. 11. Shear buckling coefficient for a simply supported rectangular plate.

3.3. Example three

A simply supported square plate with a central circular cut-out subjected to uniaxial compression is studied. A quarter of
the plate (see Fig. 12) is chosen as the computational model due to its double symmetry. The geometric transformation of the
plate has already been discussed in Section 2.1.
Symmetric boundary conditions are enforced on the vertical side AF and the horizontal side CD, i.e.
u ¼ ux ¼ 0; for the vertical side AF ð59Þ
and
v ¼ uy ¼ 0; for the horizontal side CD: ð60Þ
5072 H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074

Nx F E Nx
F E Nx

A (2)
B
C D L/2
a L
A B (1)
y a

C D
o x
L
L/2

Fig. 12. A quarter of a simply supported square plate with central circular cut-out under uni-axial compression.

Fig. 13. Buckling load coefficient for thin square plate with a central cut-out.

A dimensionless eigenbuckling load coefficient is defined as

ðNx Þcr L2
k3 ¼ : ð61Þ
p2 D
In addition to the quadrature element solution which is computed with N = 10 for all cases, the ANSYS results using the
Shell181 element as well as the p-type finite element predictions (p = 9) are obtained for comparison.
Fig. 13 shows the solutions for thin plate h/L = 0.001. Excellent agreement is reached for thin plates among the three
numerical solutions with the change of the cut-out size a/L. The buckling coefficient decreases and the corresponding buck-
ling mode remains out-of-plane with the increase of circular cut-out size. For moderately thick plates, however, the buckling
mode switches from an out-of-plane shape into an in-plane shape, giving rise to precipitation of the buckling load curves
when a/L is increased to 0.34 (see Fig. 14). The quadrature element solution agrees very well with the p-type finite element
prediction for various sizes of the cut-out. In addition, it is observed that the ANSYS solution is a little higher than those of
the QEM and the p-FEM when the plate is buckled into an out-of-plane shape. Before the mode switching point is reached,
the larger the cut-out, the larger the disparity between the ANSYS prediction and the quadrature element solution. It is be-
lieved that this is accounted for by the inclusion of rotational displacements in Eq. (15), being usually neglected in finite ele-
ment analysis. For plate with large cut-out (a/L > 0.34), excellent agreement is reached among the quadrature element
solution, the p-type finite element solution as well as the ANSYS prediction since the buckling mode is in-plane and the rota-
tional displacements in Eq. (15) are irrelevant.
H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074 5073

Fig. 14. Buckling load coefficient for moderately thick square plate with a central cut-out.

4. Concluding remarks

It has been shown that the weak form quadrature element method is an effective way to evaluate to the buckling loads of
first order shear deformable plates. Three typical examples have been studied to verify the method. Although the quadrature
element solution achieves virtually the same accuracy as either the p-type finite element solution or h-type finite element
solution, it enjoys high computational efficiency and modeling advantage. In practice, the order of approximation for a p-
type finite element is usually chosen to be less than 10 and further division is needed even though the domain of interest
is regular. In contrast, the quadrature element solution has virtually no limitation on the order of approximation. Therefore,
it can be counted on to deal with problem divided into large size quadrilateral elements. In addition, it has been shown in
present investigation that nonlinear strain terms associated with rotational displacements should be included in buckling
analysis of moderately thick plates since omission of these terms may result in significant overestimation of the critical load.
It is believed that the weak form quadrature element method may act as a competitive numerical tool for analysis of mod-
erately thick plates, due to its pronounced succinctness and efficacy.

Acknowledgement

This study was undertaken under the support of the National Natural Science Foundation of China (No. 50778104) and
State 863 Project (No. 2009AA04Z401).

References

[1] E. Reissner, The effects of transverse shear deformation on the bending of elastic plates, J. Appl. Mech. 67 (1945) 69–772.
[2] R.D. Mindlin, Influence of rotatory inertia and shear in flexural motion of isotropic elastic plates, J. Appl. Mech. 18 (1951) 1031–1036.
[3] H. Zhong, T. Yu, Flexural vibration analysis of an eccentric annular Mindlin plate, Archive Appl. Mech. 77 (2007) 185–195.
[4] 4.H. Zhong, T. Yu, A weak form quadrature element method for plane elasticity problems, Appl. Math. Modell 33 (2009) 3801–3814.
[5] Y. Mo, L. Ou, H. Zhong, Vibration analysis of Timoshenko beams on a nonlinear elastic foundation, Tsinghua Sci. Technol. 14 (3) (2009) 322–326.
[6] H. Zhong, M. Gao, Quadrature element analysis of planar frameworks, Archive Appl. Mech., doi:10.1007/s00419-009-0388-9.
[7] H. Zhong, R. Zhang, H. Yu, Buckling Analysis of planar frameworks using the quadrature element method, Int. J. Struct. Stability. Dyn, in press.
[8] H. Zhong, Y. Wang, Weak form quadrature element analysis of Bickford Beams, Eur. J.Mech./Solid., doi:10.1016/j.euromechsol.2010.03.012.
[9] A.G. Striz, W.L. Chen, C.W. Bert, Static analysis of structures by the quadrature element method (QEM), Int. J. Solids Struct. 31 (20) (1994) 2807–2818.
[10] A.G. Striz, W.L. Chen, C.W. Bert, Free vibration of plates by the high accuracy quadrature element method, J. Sound Vib. 202 (1997) 689–702.
[11] B.A. Szabó, I. Babuška, Finite Element Analysis, John Wiley & Sons, New York, 1991.
[12] W.J. Gordon, Blending function methods of bivariate and multivariate interpolation and approximation, SIAM J. Numer. Anal. 8 (1971) 158–177.
[13] P.Z. Bazant, L. Cedolin, Stability of Structures: Elastic, Inelastic, Fracture and Damage Theories, Oxford University Press, New York, 1991.
[14] H.L. Langhaar, Energy Methods in Applied Mechanics, John Wiley & Sons, New York, 1962.
[15] P.I. Davis, P. Rabinowitz, Methods of Numerical Integration, second ed., Academic Press, Orlando, 1984.
[16] R.E. Bellman, J. Casti, Differential quadrature and long term integration, J. Math. Anal. Appl. 34 (1971) 235–238.
[17] C.W. Bert, M. Malik, Differential quadrature method in computational mechanics: a review, Appl. Mech. Rev. 49 (1996) 1–27.
[18] C. Shu, Differential Quadrature and Its Application in Engineering, Springer-Verlag, London, 2000.
5074 H. Zhong et al. / Applied Mathematical Modelling 35 (2011) 5059–5074

[19] S.P. Timoshenko, J.M. Gere, Theory of Elastic Stability, McGraw-Hill, Singapore, 1988.
[20] ANSYS, Swanson Analysis System, US, 2003.
[21] I. Babuska, M. Griebel, J. Pitkranta, The problem of selecting the shape functions for a p-type finite element, Int. J. Numer. Methods Eng. 28 (1989)
1891–1908.

You might also like