Advances in The Formulation of The Rotation-Free Basic Shell Triangle
Advances in The Formulation of The Rotation-Free Basic Shell Triangle
Received 31 October 2003; received in revised form 1 March 2004; accepted 20 July 2004
Abstract
A family of rotation-free three node triangular shell elements is presented. The simplest element of the family is
based on an assumed constant curvature field expressed in terms of the nodal deflections of a patch of four elements
and a constant membrane field computed from the standard linear interpolation of the displacements within each tri-
angle. An enhanced version of the element is obtained by using a quadratic interpolation of the geometry in terms of the
six patch nodes. This allows to compute an assumed linear membrane strain field which improves the in-plane behav-
iour of the original element. A simple and economic version of the element using a single integration point is presented.
The efficiency of the different rotation-free shell triangles is demonstrated in many examples of application including
linear and non-linear analysis of shells under static and dynamic loads, the inflation and de-inflation of membranes
and a sheet stamping problem.
2005 Elsevier B.V. All rights reserved.
1. Introduction
Triangular shell elements are very useful for the solution of large scale shell problems such as those
occurring in many practical engineering situations. Typical examples are the analysis of shell roofs under
static and dynamic loads, sheet stamping processes, vehicle dynamics and crash-worthiness situations.
Many of these problems involve high geometrical and material non-linearities and time changing frictional
*
Corresponding author. Tel.: +34 93 205 7016; fax: +34 93 401 6517.
E-mail address: [email protected] (E. Oñate).
0045-7825/$ - see front matter 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.cma.2004.07.039
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2407
contact conditions. These difficulties are usually increased by the need of discretizing complex geometrical
shapes. Here the use of shell triangles and non-structured meshes becomes a critical necessity. Despite re-
cent advances in the field [1–6] there are not so many simple shell triangles which are capable of accurately
modelling the deformation of a shell structure under arbitrary loading conditions.
A promising line to derive simple shell triangles is to use the nodal displacements as the only unknowns
for describing the shell kinematics. This idea goes back to the original attempts to solve thin plate bending
problems using finite difference schemes with the deflection as the only nodal variable [7–9].
In past years some authors have derived a number of thin plate and shell triangular elements free of rota-
tional degrees of freedom (d.o.f.) based on Kirchhoffs theory [10–26]. In essence all methods attempt to
express the curvatures field over an element in terms of the displacements of a collection of nodes belonging
to a patch of adjacent elements. Oñate and Cervera [14] proposed a general procedure of this kind combin-
ing finite element and finite volume concepts for deriving thin plate triangles and quadrilaterals with the
deflection as the only nodal variable and presented a simple and competitive rotation-free three d.o.f. tri-
angular element termed BPT (for Basic Plate Triangle). These ideas were extended and formalized in [20] to
derive a number of rotation-free thin plate and shell triangles. The basic ingredients of the method are a
mixed Hu–Washizu formulation, a standard discretization into three node triangles, a linear finite element
interpolation of the displacement field within each triangle and a finite volume type approach for comput-
ing constant curvature and bending moment fields within appropriate non-overlapping control domains.
The so called ‘‘cell-centered’’ and ‘‘cell-vertex’’ triangular domains yield different families of rotation-free
plate and shell triangles. Both the BPT plate element and its extension to shell analysis (termed BST for
basic shell triangle) can be derived from the cell-centered formulation. Here the ‘‘control domain’’ is an
individual triangle. The constant curvatures field within a triangle is computed in terms of the
displacements of the six nodes belonging to the four elements patch formed by the chosen triangle and
the three adjacent triangles. The cell-vertex approach yields a different family of rotation-free plate and
shell triangles. Details of the derivation of both rotation-free triangular shell element families can be found
in [20].
An extension of the BST element to the non-linear analysis of shells was implemented in an explicit dy-
namic code by Oñate et al. [25] using an updated Lagrangian formulation and a hypo-elastic constitutive
model. Excellent numerical results were obtained for non-linear dynamics of shells involving frictional con-
tact situations and sheet stamping problems [17–19,25].
A large strain formulation for the BST element using a total Lagrangian description was presented by
Flores and Oñate [23]. A recent extension of this formulation is based on a quadratic interpolation of
the geometry of the patch formed by the BST element and the three adjacent triangles [26]. This yields a
linear displacement gradient field over the element from which linear membrane strains and constant cur-
vatures can be computed within the BST element.
In this paper the formulation of the BST element is revisited using an ‘‘assumed strain’’ approach. The
content of the paper is the following. First some basic concepts of the formulation of the original BST ele-
ment using an assumed constant curvature field are given. Next, the basic equations of the non-linear thin
shell theory chosen based on a total Lagrangian description are presented. Then the non-linear formulation
of the BST element is presented. This is based on an assumed constant membrane field derived from the
linear displacement interpolation and an assumed constant curvature field expressed in terms of the dis-
placements of the nodes of the four element patch using a finite volume type approach. An enhanced ver-
sion of the BST element is derived using an assumed linear field for the membrane strains and an assumed
constant curvature field. Both assumed fields are obtained from the quadratic interpolation of the patch
geometry following the ideas presented in [26]. Details of the derivation of the tangent stiffness matrix
needed for a quasi-static implicit solution are given for both the BST and EBST elements. An efficient ver-
sion of the EBST element using one single quadrature point for integration of the tangent matrix is pre-
sented. An explicit scheme adequate for dynamic analysis is briefly described.
2408 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
The efficiency and accuracy of the standard and enhanced versions of the BST element is validated in a
number of examples of application including linear and non-linear analysis of shells under static and dy-
namic loads, the inflation and de-inflation of membranes and a sheet stamping problem.
2. Formulation of the basic plate triangle using an assumed constant curvature field
Let us consider a patch of four plate three node triangles (Fig. 1). The nodes 1, 2, and 3 in the main
central triangle (M) are marked with circles while the external nodes in the patch (nodes 4, 5 and 6) are
marked with squares. Mid-side points in the central triangle are also marked with smaller squares. Kirch-
hoffs thin plate theory will be assumed to hold. The deflection is linearly interpolated within each three
node triangle in the standard finite element manner as
X
3
w¼ Lei wei ; ð1Þ
i¼1
where Lei are the linear shape functions (area coordinates) of the three node triangle, wei are nodal deflec-
tions and superindex e denotes element values.
The curvatures within the central triangle can be expressed in terms of a constant assumed curvatures
field as
8 9
< jxx >
> =
j ¼ jyy ¼ j ^; ð2Þ
>
: >
;
jxy
1
3 1 2
2
M
3
2 3
1
5
6
Fig. 1. Patch of three node triangular elements including the central triangle (M) and three adjacent triangles (1, 2 and 3).
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2409
2 38 ow 9
I nx 0 > >
<
>
>
=
1 6 7 ox
j¼ 4 0 ny 5 dC; ð4Þ
AM >
> ow >
>
CM ny nx : ;
oy
where CM is the boundary of the central triangle and n = (nx, ny) is the boundary normal. Eq. (4) defines the
assumed constant curvature field within the central triangle in terms of the deflection gradient along the
sides of the triangle. Eq. (4) can be found to be equivalent to the standard conservation laws used in finite
volume procedures as described in [27,28].
The computation of the line integral in Eq. (4) poses a difficulty as the deflection gradient is discontin-
uous along the element sides. A simple method to overcome this problem is to compute the deflection gra-
dient at the element sides as the average values of the gradient contributed by the two triangles sharing the
side [20,28]. Following this idea the constant curvature field with the element is computed as
2 j 3M 2 M 3
nx 0 Lj;x 0
1 X lj 6 X6
3 M 3
7 M M
j¼ 4 0 njy 7
5 rLM M j j
i wi þ rLi wi ¼
6 0 LM
4
7 j j p
j;y 5 rLi wi þ rLi wi ¼ Bb w ;
AM j¼1 2
njy njx j¼1
LMj;y LMj;x
ð5Þ
p T
where w = [w1, w2, w3, w4, w5, w6] is the deflection vector of the six nodes in the patch. In Eq. (5) the sum
extends over the three sides of the central element M, lMj are the lengths of the element sides and superin-
dexes M and j refer to the central triangle and to each of the adjacent elements, respectively. The standard
sum convention for repeated indexes is used. Note that triangular area coordinates satisfy
" M# " #
M
Li;x lM
i
nix
rLi ¼ ¼ : ð6Þ
LMi;y
2AM niy
Note also that the constant curvature field is expressed in terms of the six nodes of the four element
patch linked to the element M. The expression of the 3 · 6 Bb matrix can be found in [14,20].
The virtual work expression is written as
Z Z Z Z
djT m dA ¼ dwq dA; ð7Þ
A A
where m is the bending moment field related to the curvatures by the standard constitutive equations
2 3
1 m 0
h3 E 6m 1 0 7
3
T
m ¼ ½M xx ; M yy ; M xy ¼ Db j; Db ¼ 6 7 ¼ h D: ð8Þ
12 ð1 m2 Þ 4 1 m 5 12
0 0
2
In Eqs. (7) and (8) h is the plate thickness, E is the Youngs modulus, m is the Poissons ratio, dj and dw
are the virtual curvatures and the virtual deflection, respectively, and q is a distributed vertical load.
Substituting the approximation for the vertical deflection and the assumed constant curvature field into
(7) leads to the standard linear system of equations
Kw ¼ f; ð9Þ
where the stiffness matrix K and the equivalent nodal force f can be found by assembly of the element con-
tributions given by
Z Z
e
K ¼ BTb Db Bb dA; ð10Þ
Ae
2410 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
8 e9
Z Z > < L1 >
=
e
f ¼ q Le2 dA: ð11Þ
: e>
Ae > ;
L3
Note that Ke is a 6 · 6 matrix, whereas fe has the same structure than for the standard linear triangle. The
explicit form of Ke and fe can be found in [14].
The resulting basic plate triangle (BPT) has one degree of freedom per node and a wider bandwidth than
the standard three node triangles as each triangular element is linked to its three neighbours through Eq.
(5).
Examples of the good performance of the BPT element for analysis of thin plates can be found in [14,20].
The extension of the BPT element to the analysis of shells yields the basic shell triangle (BST) [20]. Different
applications of the BST element to linear and non-linear analysis of shells are reported in [14,17–
20,23,25,26].
The ideas used to derive the BPT element will now be extended to derive two families of basic shell tri-
angles using a total Lagrangian description.
A summary of the most relevant hypothesis related to the kinematic behaviour of a thin shell are pre-
sented. Further details may be found in the wide literature dedicated to this field [8,9].
Consider a shell with undeformed middle surface occupying the domain X0 in R3 with a boundary C0. At
each point of the middle surface a thickness h0 is defined. The positions x0 and x of a point in the unde-
formed and the deformed configurations can be, respectively, written as a function of the coordinates of
the middle surface u and the normal t3 at the point as
where n1, n2 are arc-length curvilinear principal coordinates defined over the middle surface of the shell and
f is the distance from the point to the middle surface in the undeformed configuration. The product fk is the
distance from the point to the middle surface measured on the deformed configuration. The parameter k
relates the thickness at the present and initial configurations as
h
k¼ : ð14Þ
h0
This approach implies a constant strain in the normal direction. Parameter k will not be considered as an
independent variable and will be computed from purely geometrical considerations (isochoric behaviour)
via a staggered iterative update. Besides this, the usual plane stress condition of thin shell theory will be
adopted.
A convective system is computed at each point as
ox
gi ðnÞ ¼ i ¼ 1; 2; 3; ð15Þ
oni
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2411
oðuðn1 ; n2 Þ þ fkt3 Þ
ga ðnÞ ¼ ¼ u0 a þ fðkt3 Þ0 a a ¼ 1; 2; ð16Þ
ona
oðuðn1 ; n2 Þ þ fkt3 Þ
g3 ðnÞ ¼ ¼ kt3 : ð17Þ
of
This can be particularized for the points on the middle surface as
aa ¼ ga ðf ¼ 0Þ ¼ u0 a ; ð18Þ
a3 ¼ g3 ðf ¼ 0Þ ¼ kt3 : ð19Þ
The covariant (first fundamental form) metric tensor of the middle surface is
aab ¼ aa ab ¼ u0 a u0 b : ð20Þ
The Green–Lagrange strain vector of the middle surface points (membrane strains) is defined as
T
em ¼ ½em11 ; em12 ; em12 ð21Þ
with
emij ¼ 12 aij a0ij : ð22Þ
The curvatures (second fundamental form) of the middle surface are obtained by
jab ¼ 12ðu0 a t30 b þ u0 b t30 a Þ ¼ t3 u0 ab ; a; b ¼ 1; 2: ð23Þ
The deformation gradient tensor is
F ¼ ½x0 1 ; x0 2 ; x0 3 ¼ ½ u0 1 þ fðkt3 Þ0 1 u0 2 þ fðkt3 Þ0 2 kt3 : ð24Þ
T 2
The product F F = U = C (where U is the right stretch tensor, and C the right Cauchy-Green deforma-
tion tensor) can be written as
2 3
a11 þ 2j11 fk a12 þ 2j12 fk 0
6 7
U2 ¼ 4 a12 þ 2j12 fk a22 þ 2j22 fk 0 5: ð25Þ
2
0 0 k
In the derivation of expression (25) the derivatives of the thickness ratio k0 a and the terms associated to f2
have been neglected.
Eq. (25) shows that U2 is not a unit tensor at the original configuration for curved surfaces (j0ij 6¼ 0). The
changes of curvature of the middle surface are computed by
This expression is useful to compute different Lagrangian strain measures. An advantage of these mea-
sures is that they are associated to material fibres, what makes it easy to take into account material aniso-
tropy. It is also useful to compute the eigendecomposition of U as
X
3
U¼ ka r a ra ; ð28Þ
a¼1
where du are virtual displacements, dem is the virtual Green–Lagrange membrane strain vector, dj are the
virtual curvatures and t are the surface loads. Other load types can be easily included into (31).
In order to treat plasticity at finite strains an adequate stress-strain pair must be used. The Hencky mea-
sures will be adopted here. The (logarithmic) strains are defined as
2 3
e11 e21 0
7 X
3
6
Eln ¼ 4 e12 e22 0 5 ¼ lnðka Þra ra : ð32Þ
a¼1
0 0 e33
Two types of material models are considered here: an elastic–plastic material associated to thin rolled
metal sheets and a hyper-elastic material for rubbers.
In the case of metals, where the elastic strains are small, the use of a logarithmic strain measure reason-
ably allows to adopt an additive decomposition of elastic and plastic components as
Eln ¼ Eeln þ Epln : ð33Þ
A constant linear relationship between the (plane) Hencky stresses and the logarithmic elastic strains is
adopted giving
T ¼ DEeln : ð34Þ
These constitutive equations are integrated using a standard return algorithm. The following Mises–Hill
[29] yield function with non-linear isotropic hardening is chosen
n
ðG þ H ÞT 211 þ ðF þ H ÞT 222 2HT 11 T 22 þ 2NT 212 ¼ r0 ðe0 þ ep Þ ; ð35Þ
where F, G, H and N define the non-isotropic shape of the yield surface and the parameters r0, e0 and n
define its size as a function of the effective plastic strain ep.
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2413
The simple Mises–Hill yield function allows, as a first approximation, to treat rolled thin metal sheets
with planar and transversal anisotropy.
For the case of rubbers, the Ogden [30] model extended to the compressible range is considered.
The material behaviour is characterized by the strain energy density per unit undeformed volume defined
as
" ! #
K 2
X N
lp ap X 3
ap 1
w ¼ ðln J Þ þ J 3 ki 3 ; ð36Þ
2 p¼1
ap i¼1
where K is the bulk modulus of the material, J is the determinant of U, N, li and ai are material parameters,
li, ai are real numbers such that liai > 0 ("i = 1, N) and N is a positive integer.
The stress measures associated to the principal logarithmic strains are denoted by bi. They can be com-
puted noting that
!
owðka Þ XN
ap
ap 1 1 1 X 3
ap
3
bi ¼ ¼ Kðln JÞ þ ki lp J ki k ð37Þ
oðln ki Þ p¼1
3 ki j¼1 j
we define now
X
3
a
ap ¼ kj p ð38Þ
j¼1
which gives
X
N
ap
3 ap 1
bi ¼ Kðln J Þ þ lp J ki ap : ð39Þ
p¼1
3
The values of bi, expressed in the principal strains directions, allow to evaluate the Hencky stresses in the
convective coordinate system as
X
3
T¼ b i ri ri : ð40Þ
i¼1
The Hencky stress tensor T can be easily particularized for the plane stress case.
We define the rotated Hencky and second Piola–Kirchhoff stress tensors as
TL ¼ RTL TRL ; ð41Þ
RL ¼ ½ r1 ; r2 ; r3 : ð43Þ
The relationship between the rotated Hencky and Piola–Kirchhoff stresses is (a, b = 1, 2)
1
½S L aa ¼ ½T L aa ;
k2a
ð44Þ
lnðka =kb Þ
½S L ab ¼1 2 ½T L ab :
2
ðka k2b Þ
2414 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
The second Piola–Kirchhoff stress vector r of Eqs. (29) and (30) can be readily extracted from the S
tensor.
4.1. Definition of the element geometry and discretization of the displacement field
The rotation-free BST element has three nodes with three displacement degrees of freedom at each node.
As for the BPT element a patch is defined by the central triangle and the three adjacent elements (Fig. 1).
This four elements patch helps to define the curvature field within the central triangle (the BST element) in
terms of the displacements of the six patch nodes.
The node-ordering in the patch is the following (see Fig. 1).
• The nodes in the main element (M) are numbered locally as 1, 2 and 3. They are defined counter-clock-
wise around the positive normal.
• The sides in the main element are numbered locally as 1, 2, and 3. They are defined by the local node
opposite to the side.
• The adjacent elements (which are part of the patch) are numbered with the number associated to the
common side.
• The extra nodes of the patch are numbered locally as 4, 5 and 6, corresponding to nodes on adjacent
elements opposite to sides 1, 2 and 3 respectively.
• The connectivities in the adjacent elements are defined beginning with the extra node as shown in
Table 1.
The following local Cartesian coordinate system is defined for the patch. In the main element the unit
vector t1 (associated to the local coordinate n1) is directed along side 3 (from node 1 to node 2), t3 (asso-
ciated to the coordinate f) is the unit normal to the plane, and finally t2 = t3 · t1 (associated to the coordi-
nate n2).
The coordinates and the displacements are linearly interpolated within each three node triangle in the
mesh in the standard manner, i.e.
X
3 X
3
u¼ Lei ui ¼ Lei u0i þ ui ; ð46Þ
i¼1 i¼1
Table 1
Element numbering and nodal connectivities of the four elements patch of Fig. 1
Element N1 N2 N3
M 1 2 3
1 4 3 2
2 5 1 3
3 6 2 1
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2415
8 9 8 9
< u1 >
> = X3 < u1 >
> =
u ¼ u2 ¼ Lei ui ; ui ¼ u2 : ð47Þ
>
: ; > i¼1 : >
> ;
u3 u3 i
In above ui and ui contain, respectively, the three coordinates and the three displacements of node i.
The Green–Lagrange membrane strains are expressed by substituting the linear displacement interpola-
tion into Eq. (22). This gives
2 3
u0 1 u0 1 1
16 7
em ¼ 4 u 0 2 u 0 2 1 5 : ð48Þ
2
2u0 1 u0 2
The membrane strain field is constant within each triangle similarly as in the standard CST element. The
variation of the membrane strains is obtained by
dem ¼ Bm dae ð49Þ
with
8 9
< u1 >
> =
Bm ¼ ½Bm1 ; Bm2 ; Bm3 ; ae ¼ u2 ð50Þ
: >
> ;
u3
and
2 3
LM T
i;1 u0 1
B mi ¼ 6 LM T 7
ð51Þ
33
4 i;2 u0 2 5:
LM T
i;1 u0 2 þ LM T
i;2 u0 1
We will assume the following constant curvature field within each element
jab ¼ j
^ab ; ð52Þ
where j
^ab is the assumed constant curvature field defined by
Z
1
^ab ¼ 0
j t3 u0 ba dA0 ; ð53Þ
AM A0M
where A0M is the area (in the original configuration) of the central element in the patch.
Substituting Eq. (53) into (52) and integrating by parts the area integral gives the curvature vector within
the element in terms of the following line integral
8 9 2 3
< j11 >
> = I n1 0
1 6 7 t3 u0 1
j¼ j22 ¼ 0 4 0 n2 5 dC; ð54Þ
>
: >
; AM C0M t3 u0 2
2j12 n2 n1
2416 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
where ni are the components (in the local system) of the normals to the element sides in the initial config-
uration C0M .
For the definition of the normal vector t3, the linear interpolation over the central element is used. In this
case the tangent plane components are
X
3
u0 a ¼ LM
i;a ui ; a ¼ 1; 2; ð55Þ
i¼1
u0 1 u0 2
t3 ¼ ¼ ku1 u2 : ð56Þ
ju0 1 u0 2 j
From these expressions it is also possible to compute in the original configuration the element area A0M ,
the outer normals (n1, n2)i at each side and the side lengths lM
i . Eq. (56) also allows to evaluate the thickness
ratio k in the deformed configuration and the actual normal t3.
In order to compute the line integral of Eq. (54) the averaging procedure described in Section 2 is used.
Hence along each side of the triangle the average value of u0 a between the main triangle and the adjacent
one is taken leading to
2 i 3
n1 0 " #
1 X M6
3
i 7
t3 12 uM i
0 1 þ u0 1
j¼ 0 l 4 0 n2 5 ; ð57Þ
AM i¼1 i i i
t3 12 uM i
0 2 þ u0 2
n2 n1
where the sum extends over the three elements adjacent to the central triangle M.
Noting that t3 uM0 a ¼ 0 in the main triangle and using (6) it can be found [23]
2 M 3
Li;1 0
X
3
6
i
7 t3 u0 1
j¼ 4 0 LM
i;2 5 : ð58Þ
i¼1 t3 ui0 2
LM
i;2
M
Li;1
This can be seen as the projection of the local derivatives of u in the adjacent triangles (ui0 a where index i
denotes values associated to the adjacent elements) over the normal to the main triangle t3. As the triangles
have a common side, t3 ui0 s ¼ 0, where ui0 s is the derivative of u along the side. Hence only the derivative of
u along the side normal (ui0 n ) has non-zero component over t3. This gives
t3 ui0 1
¼ t3 ui0 n ni : ð59Þ
t3 ui0 2
An alternative form to express the curvatures, which is useful when their variations are needed, is to de-
fine the vectors
X3
1 M k
hij ¼ Lk;i u0 j þ LM uk
k;j 0i : ð60Þ
k¼1
2
This gives
jij ¼ hij t3 : ð61Þ
The last expression allows to interpret the curvatures as the projections of the vectors hij over the normal
of the central element. The variation of the curvatures can be obtained as
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2417
8 2 3 9
>
> 2 M 3 2 3 LM q111 þ LM q211 >
>
>
> Li;1 0 6 i;1 i;2
7 >
>
X >
3 <6 X 3 L i
t du i
6 7 >
=
7 6 j;1 3 j
7 6 M 1 7
dj ¼ 6 0 M 7
Li;2 5 4 5 2 6 i;1 22q þ M 2
q du M
ð62Þ
L L i;2 22 7 3 t ;
>4
i¼1 > i i 6 7 i
>
>
>
> J ¼1 Lj;2 t3 duj 4 5 >
>
> M
: Li;2 LM >
;
i;1 LM 1 M 2
i;1 q12 þ Li;2 q12
where the projections of the vectors hij over the contravariant base vectors ua have been included
with
u1 ¼ ku0 2 t3 ; ð64Þ
u2 ¼ ku0 1 t3 ; ð65Þ
In above expressions superindexes in and Lkj dukj
refer to element numbers in the patch whereas subscripts
denote node numbers of each element in the patch. As usual the superindex M denotes values in the central
triangle (Fig. 1). Note that as expected the curvatures (and their variations) in the central element are a
function of the nodal displacements of the six nodes in the four elements patch. Note also the isochoric
approach
h A0
k¼ 0
¼ M: ð66Þ
h AM
The derivation of Eq. (62) can be found in [26]. This equation can be rewritten in the form
dj ¼ Bb dap ; ð67Þ
where
dap ¼ duT ; duT ; duT ; duT ; duT ; duT T ð68Þ
181 1 2 3 4 5 6
An enhanced version of the BST element (termed EBST) has been recently proposed by Flores and
Oñate [26]. The main features of the element formulation are the following:
1. The geometry of the patch formed by the central element and the three adjacent elements is quadratically
interpolated from the position of the six nodes in the patch.
2. The membrane strains are assumed to vary linearly within the central triangle and are expressed in terms
of the (continuous) values of the deformation gradient at the mid-side points of the triangle.
3. The assumed constant curvature field within the central triangle is obtained by expression (54) using now
twice the values of the (continuous) deformation gradient at the mid-side points.
2418 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
As mentioned above a quadratic approximation of the geometry of the four elements patch is chosen
using the position of the six nodes in the patch. It is useful to define the patch in the isoparametric space
using the nodal positions given in the Table 2 (see also Fig. 2).
The quadratic interpolation is defined by
X6
u¼ N i ui ð70Þ
i¼1
with (f = 1 n g)
f
N 1 ¼ f þ ng N 4 ¼ ðf 1Þ;
2
n ð71Þ
N 2 ¼ n þ gf N 5 ¼ ðn 1Þ;
2
g
N 3 ¼ g þ fn N 6 ¼ ðg 1Þ:
2
This interpolation allows to compute the displacement gradients at selected points in order to use an as-
sumed strain approach. The computation of the gradients is performed at the mid-side points of the central
element of the patch denoted by G1, G2 and G3 in Fig. 2. This choice has the following advantages.
• Gradients at the three mid-side points depend only on the nodes belonging to the two elements adjacent
to each side. This can be easily verified by sampling the derivatives of the shape functions at each mid-
side point.
• When gradients are computed at the common mid-side point of two adjacent elements, the same values
are obtained, as the coordinates of the same four points are used. This in practice means that the gra-
dients at the mid-side points are independent of the element where they are computed. A side-oriented
implementation of the finite element will therefore lead to a unique evaluation of the gradients per side.
Table 2
Isoparametric coordinates of the six nodes in the patch of Fig. 2
1 2 3 4 5 6
n 0 1 0 1 1 1
g 0 0 1 1 1 1
η
5 3 4
..
G2 G1
. 1
G3
2
ξ
The Cartesian derivatives of the shape functions are computed at the original configuration by the stan-
dard expression
N i;1 N i;n
¼ J1 ; ð72Þ
N i;2 N i;g
where the Jacobian matrix at the original configuration is
" 0 #
u0 n t1 u00 g t1
J¼ : ð73Þ
u00 n t2 u00 g t2
The deformation gradients on the middle surface, associated to an arbitrary spatial Cartesian system and
to the material Cartesian system defined on the middle surface are related by
½u0 1 ; u0 2 ¼ ½u0 n ; u0 g J1 : ð74Þ
The membrane strains within the central triangle are now obtained using a linear assumed strain field ^em .
If, for example, Green Lagrange strains are used, i.e.
em ¼ ^em ð75Þ
with
X
3
^em ¼ ð1 2fÞe1m þ ð1 2nÞe2m þ ð1 2gÞe3m ¼ N i eim ; ð76Þ
i¼1
where eim are the membrane strains computed at the three mid-side points Gi (i = 1, 2, 3 see Fig. 2). In Eq.
(76) N 1 ¼ ð1 2fÞ, etc.
The gradient at each mid-side point is computed from the quadratic interpolation (70):
" #
X
3
i
ðu0 a ÞGi ¼ u0 a ¼ N j;a uj þ N iiþ3;a uiþ3 ; a ¼ 1; 2; i ¼ 1; 2; 3:
i
ð77Þ
j¼1
Substituting Eq. (22) into (76) and using Eq. (20) gives the membrane strain vector as
8 i 9
>
< u0 1 ui0 1 1 >
=
X13
em ¼ N i ui0 2 ui0 2 1 ð78Þ
i¼1
2 > : >
;
2ui0 1 ui0 2
and the virtual membrane strains as
8 9
>
< ui0 1 dui0 1 >
=
X3
dem ¼ Ni ui2 dui0 2 : ð79Þ
i¼1
>
: i i i
>
i ;
du0 1 u0 2 þ u0 1 du2
We note that the gradient at each mid-side point Gi depends only on the coordinates of the three nodes
of the central triangle and on those of an additional node in the patch, associated to the side i where the
gradient is computed.
Combining Eqs. (79) and (77) gives
Differently from the original BST element the membrane strains within the EBST element are now a
function of the displacements of the six patch nodes.
The constant curvature field assumed for the BST element is chosen again here. The numerical evalua-
tion of the line integral in Eq. (54) results in a sum over the integration points at the element boundary
which are, in fact, the same points used for evaluating the gradients when computing the membrane strains.
As one integration point is used over each side, it is not necessary to distinguish between sides (i) and inte-
gration points (Gi). In this way the curvatures can be computed by
2 M 3
Li;1 0
X 3
6 0 LM 7 t3 u0 1
i
j¼2 4 i;2 5 : ð81Þ
i¼1 M M
t3 ui0 2
Li;2 Li;1
In the standard BST element [20,23] the gradient ui0 a is computed as the average of the linear approxi-
mations over the two adjacent elements (see Section 4.3). In the enhanced version, the gradient is evaluated
at each side Gi from the quadratic interpolation
2 3
u1
" # " i #
ui01 N 1;1 N i2;1 N i3;1 N iiþ3;1 6 7
6 u2 7
¼ 6 7: ð82Þ
ui02 N i1;2 N i2;2 N i3;2 N iiþ3;2 4 u3 5
uiþ3
Note again than at each side the gradients depend only on the positions of the three nodes of the central
triangle and of an extra node (i + 3), associated precisely to the side (Gi) where the gradient is computed.
Direction t3 in Eq. (82) can be seen as a reference direction. If a different direction than that given by Eq.
(56) is chosen at an angle h with the former, this has an influence of order h2 in the projection. This justifies
Eq. (56) for the definition of t3 as a function exclusively of the three nodes of the central triangle, instead of
using the 6-node isoparametric interpolation.
The variation of the curvatures can be obtained as
2 M 3
Li;1 0 ( " i # " i #)
X63 X 3 N j;1 t3 duj N iþ3;1 ðt3 duiþ3 Þ
M 7
dj ¼ 2 4 0 Li;2 5 þ
i¼1 M M i¼1 N ij;2 t3 duj N iiþ3;2 ðt3 duiþ3 Þ
Li;2 Li;1
2 3
LM q1
i;1 11 þ L M 2
q
i;2 11
X 3 6
6 M 1 7
7
6 L q þ LM q2 7ðt3 dui Þ ¼ Bb dap ; ð83Þ
6 i;1 22 i;2 22 7
i¼1 4 5
LM 1 M 2
i;1 q12 þ Li;2 q12
where the definitions (61) and (63) still hold but with the new definition of hij given by [26]
3
X
hij ¼ LM k M k
k;i u0 j þ Lk;j u0 i : ð84Þ
k¼1
In Eq. (83)
Bb ¼ ½Bb1 ; Bb2 ; . . . ; Bb6 : ð85Þ
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2421
The expression of the curvature matrix Bb is given in the Appendix A. Details of the derivation of Eq.
(83) can be found in [26].
A simplified and yet very effective version of the EBST element can be obtained by using one point quad-
rature for the computation of all the element integrals. This element is termed EBST1. Note that this only
affects the membrane stiffness matrices and it is equivalent to using a assumed constant membrane strain
field defined by an average of the metric tensors computed at each side.
Numerical experiments have shown that both the EBST and the EBST1 elements are free of spurious
energy modes.
6. Boundary conditions
Elements at the domain boundary, where an adjacent element does not exist, deserve a special attention.
The treatment of essential boundary conditions associated to translational constraints is straightforward, as
they are the natural degrees of freedom of the element. The conditions associated to the normal vector are
crucial in the bending formulation. For clamped sides or symmetry planes, the normal vector t3 must be
kept fixed (clamped case), or constrained to move in the plane of symmetry (symmetry case). The former
case can be seen as a special case of the latter, so we will consider symmetry planes only. This restriction can
be imposed through the definition of the tangent plane at the boundary, including the normal to the plane
of symmetry u00 n that does not change during the process.
The tangent plane at the boundary (mid-side point) is expressed in terms of two orthogonal unit vectors
referred to a local-to-the-boundary Cartesian system (see Fig. 3) defined as
0
0s ;
u0 n ; u ð86Þ
where vector u00 n is fixed during the process while direction u
0 s emerges from the intersection of the sym-
metry plane with the plane defined by the central element (M). The plane (gradient) defined by the central
element in the selected original convective Cartesian system (t1, t2) is
M M
u0 1 ; u0 2 ð87Þ
the intersection line (side i) of this plane with the plane of symmetry can be written in terms of the position
of the nodes that define the side (j and k) and the original length of the side lM i , i.e.
symmetry
i
t0i3 plane t3
0 0
ϕ ’n ι =ϕ ’n
n= 0
ϕ ’n j
j
= ϕ ’s
i Μ
s ϕ ’s ϕ ’1
M t1 k Μ
k Z
ϕ ’2
t2 k
j i Y i
original X
deformed
Fig. 3. Local Cartesian system for the treatment of symmetry boundary conditions.
2422 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
1
ui0 s ¼ ðuk uj Þ: ð88Þ
lM
i
That together with the outer normal to the side ni = [n1, n2]T = [n Æ t1, n Æ t2]T (resolved in the selected ori-
ginal convective Cartesian system) leads to
" #
uiT
01 n1 n2 uiT 0n
¼ ; ð89Þ
uiT
02 n2 n1 uiT
0s
where noting that k is the determinant of the gradient, the normal component of the gradient ui0 n can be
approximated by
u00 n
ui0 n ¼ : ð90Þ
kjui0 s j
In this way the contribution of the gradient at side i to vectors hab (Eqs. (60) and (84)) results in
2 T 3i 2 M 3 2 M 3
h11 Li;1 0 " # Li;1 0 " #" iT #
6 T 7 6 7 uiT
01 6 7 n1 n2 u0 n
6 h 7 ¼ 26 0 LM 7 6 M 7
¼ 24 0 Li;2 5 : ð91Þ
4 22 5 4 i;2 5 iT
u0 2 n2 n1 uiT
0s
T
2h12 LM
i;2 LM
i;1 LM
i;2 LM
i;1
For the computation of the curvature variations, the contribution from the gradient at side i is now (see
[26])
2 T 3i 2 M 3 2 3
h11 Li;1 0 0
6 T 7 6 0 LM 7 n1 n2 4 1 5;
d4 h22 5 ¼ 24 i;2 5 ð92aÞ
n 2 n 1 ½duk duj T
T
2h12 M M
Li;2 Li;1 L o
2 3
LM
i;1 n2
2 6
LM 7 T
¼ 4 i;2 n1 5 duk duj ; ð92bÞ
lM
i
LM
i;1 n1 LM
i;2 n2
where the influence of variations in the length of vector u0 n has been neglected.
For a simple supported (hinged) side, the problem is not completely defined. The simplest choice is to
neglect the contribution to the side rotations from the adjacent element missing in the patch in the evalu-
ation of the curvatures via Eq. (54) [20,23]. This is equivalent to assume that the gradient at the side is equal
to the gradient in the central element, i.e.
ui0 1 ; ui0 2 ¼ uM M
0 1 ; u0 2 : ð93Þ
More precise changes can be however introduced to account for the different natural boundary condi-
tions. One may assume that the curvature normal to the side is zero, and consider a contribution of the
missing side to introduce this constraint. As the change of curvature parallel to the side is also zero along
the hinged side, this obviously leads to zero curvatures in both directions. Denoting the contribution to the
curvatures of the interior sides (j and k) by
2 3jk
j11
6 7
4 j22 5 :
j12
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2423
It can be easily shown that in order to set the normal curvature to zero the contribution of the simple
supported side (i) should be
2 3i 2 4 2 2 3 32 3jk
j11 ðn1 Þ ðn1 Þ ðn2 Þ ðn1 Þ n2 j11
6 7 6 76 7
4 j22 5 ¼ 4 ðn1 Þ2 ðn2 Þ2 ðn2 Þ
4 3
n1 ðn2 Þ 54 j22 5 : ð94Þ
j12 3 3 2 2 j12
2ðn1 Þ n2 2n1 ðn2 Þ 2ðn1 Þ ðn2 Þ
For the case of a triangle with two sides associated to hinged sides, the normal curvatures to both sides
must be zero. Denoting by ni and nj the normal to the sides, and by mi and mj the dual base (associated to
the base ni nj), the contribution from the hinged sides (i and j) can be written as a function of the con-
tribution of the only interior side (k):
2 3ij 2 3 2 3k
j11 mi1 mj1 j11
6 7 6 7 i j 6 7
4 j22 5 ¼ 4 mi2 mj2 5 2n1 n1 2ni2 nj2 ni1 nj2 þ ni2 nj1 4 j22 5 : ð95Þ
j12 mi1 mj2 þ mi2 mj1 j12
For a free edge the same approximation can be used but due to Poissons effect this will lead to some
error. The curvature variations of these contributions can be easily computed.
For the membrane formulation of element EBST, the gradient at the mid-side point of the boundary is
assumed equal to the gradient of the main triangle.
For a step n the configuration un and the plastic strains enp are known. The configuration un is obtained
by adding the total displacements to the original configuration un = u0 + un. The stresses are computed at
each triangle using a single sampling (integration) point at the center and NL integration points (layers)
through the thickness. The plane stress state condition of the classical thin shell theory is assumed, so that
for every layer three stress components are computed, (r11, r22, and r12) referred to the local Cartesian
system.
The computation of the incremental stresses is as follows:
1. Evaluate the incremental displacements: Dun ¼ KnT rn where KT is the tangent stiffness matrix and r is the
residual force vector defined by for each element
Z Z Z Z
T
rei ¼ Li t dA Bmi rm þ BTbi rb dA: ð96Þ
A A
The expression of the tangent stiffness matrix for the element is given below. Details of the derivation
can be found in [23,26].
2. Generate the actual configuration un+1 = un + Dun.
3. Compute the metric tensor anþ1 nþ1
ab and the curvatures jab . Then at each layer k compute the (approxi-
mate) right Cauchy-Green tensor (27)
Cnþ1
k ¼ anþ1 þ zk vnþ1 : ð97Þ
4. Compute the total (32) and elastic (33) deformations at each layer k
eknþ1 ¼ 12 ln Cknþ1 ;
nþ1 n
ð98Þ
½ee k ¼ eknþ1 ½ep k :
2424 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
As usual the tangent stiffness matrix is split into material and geometric components. The material tan-
gent stiffness matrix is computed by the integral
Z Z
KM ¼ BT Dep B dA; ð102Þ
A0M
where B = Bm + Bb includes:
Matrix Dep is the elastic–plastic tangent constitutive matrix integrated through the thickness.
A three point quadrature is used for integrating the stiffness terms BTm Dep Bm (recall that for the EBST
element the membrane strains vary linearly within the element) whereas one point quadrature is chosen
for the rest of the terms in KM.
For the EBST element the membrane part is computed as the sum of the contributions of the three sides,
i.e.
( " #" k # )
T G AM X3 X 6 X6
k k
N k11 N k12 N j;1
du Km Du ¼ dui N i;1 N i;2 k k
Duj ; ð105Þ
3 k¼1 i¼1 j¼1 N 21 N 22 N kj;2
where Nij = rmij are the axial forces defined in Eq. (29).
The geometric stiffness associated to bending moments is much more involved and can be found in [26].
Numerical experiments have shown that the bending part of the geometric stiffness is not so important and
can be disregarded in the iterative process.
Again three and one point quadratures are used for computing the membrane and bending contributions
to the geometric stiffness matrix. We note that for elastic–plastic problems a uniform one point quadrature
has been chosen for integrating both the membrane and bending stiffness matrices.
For simulations including large non-linearities, such as frictional contact conditions on complex geo-
metries or large instabilities in membranes, convergence is difficult to achieve with implicit schemes. In
those cases an explicit solution algorithm is typically most advantageous. This scheme provides the solution
for dynamic problems and also for static problems if an adequate damping is chosen.
The dynamic equations of motion to solve are of the form
rðuÞ þ Cu_ þ M€
u ¼ 0; ð106Þ
where M is the mass matrix, C is the damping matrix and the dot means the time derivative. The solution is
performed using the central difference method. To make the method competitive a diagonal (lumped) M ma-
trix is typically used and C is taken proportional to M. As usual, mass lumping is performed by assigning
one third of the triangular element mass to each node in the central element.
The explicit solution scheme can be summarized as follows. At each time step n where displacements
have been computed:
1. Compute the internal forces rn. This follows the same steps (2–8) described for the implicit scheme in the
previous section.
2. Compute the accelerations at time tn
n
un ¼ M1
€ d r Cu_ n1=2 ;
where Md is the diagonal (lumped) mass matrix.
3. Compute the velocities at time tn+1/2
u_ nþ1=2 ¼ u_ n1=2 €
un dt:
4. Compute the displacements at time tn+1
unþ1 ¼ un þ u_ nþ1=2 dt:
5. Update the shell geometry.
6. Check frictional contact conditions.
Further details of the implementation of the standard BST element within an explicit solution scheme
can be found in [25].
2426 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
9. Examples
In this section several examples are presented to show the good performance of the rotation-free shell
elements BST, EBST and EBST1. The first five static examples are solved using an implicit code. The rest
of the examples are computed using the explicit dynamic scheme. For the explicit scheme the EBST element
is always integrated using one integration point per element (EBST1 version) although not indicated.
The three elements considered (BST, EBST and EBST1) satisfy the membrane patch test defined in Fig.
4. A uniform axial tensile stress is obtained in all cases.
The element bending formulation does not allow to apply external bending moments (there are not rota-
tional DOFs). Hence it is not possible to analyse a patch of elements under loads leading to a uniform
bending moment. A uniform torsion can be considered if a point load is applied at the corner of a rectan-
gular plate with two consecutive free sides and two simple supported sides. Fig. 5 shows three patches lead-
ing to correct results both in displacements and stresses. All three patches are structured meshes. When the
central node in the third patch is shifted from the center, the results obtained with the EBST and EBST1
elements are not correct. This however does not seems to preclude the excellent performance of these ele-
ments, as proved in the rest of the examples analyzed. On the other hand, the BST element gives correct
results in all torsion patch tests if natural boundary conditions are imposed in the formulation. If this is
not the case, incorrect results are obtained even with structured meshes.
This example is used to assess the membrane performance of the EBST and EBST1 elements and to com-
pare it with the standard linear triangle (constant strain) and the quadratic triangle (linear strain). This
example involves important shear energy and was proposed to assess the distortion capability of elements.
1/4
1/2
1/4
Free
Fr
ee
ss
ss
F 25
16
Vertex displacement
20
15
QUAD4
Clamped
LST
44
E=1 10 CST
EBST
ν=0.33 EBST1
5
48
0 20 40 60 80
(a) (b) Number of nodes
Fig. 6. Cooks membrane problem: (a) geometry, (b) results.
Fig. 6a shows the geometry and the applied load. Fig. 6b plots the vertical displacement of the upper vertex
as a function of the number of nodes in the mesh. Results obtained with other isoparametric elements have
also been plotted for comparison. They include the constant strain triangle (CST), the bilinear quadrilateral
(QUAD4) and the linear strain triangle (LST) [4].
From the plot shown it can be seen that the enhanced element with three integration points (EBST) gives
values slightly better that the constant strain triangle for the coarsest mesh (only two elements). However,
when the mesh is refined a performance similar to the linear strain triangle is obtained that is dramatically
superior than the former. On the other hand, if a one point quadrature is used (EBST1) the convergence in
the reported displacement is notably better than for the rest of the elements.
In this example an effective membrane interpolation is of primary importance. The geometry is a cylin-
drical roof supported by a rigid diaphragm at both ends and it is loaded by a uniform dead weight (see
Fig. 7a). Only one quarter of the structure is modelled due to symmetry conditions. Unstructured and struc-
tured meshes are considered. In the latter case two orientations are possible (Fig. 7a shows orientation B).
Tables 3–5 present the normalized vertical displacements at the crown (point A) and at the midpoint of
the free side (point B) for the two orientations of the structured meshes and for the non-structured mesh.
Values used for normalization are uA = 0.5407 and uB = 3.610 that are quoted in reference [31].
Plots in Fig. 7b show the normalized displacement of point-B for structured meshes as a function of the
number of degrees of freedom for each case studied. An excellent convergence for the EBST element can be
seen. The version with only one integration point (EBST1) presents a behavior a little more flexible and
converges from above for structured meshes. Table 5 shows that both the EBST and the EBST1 elements
have an excellent behavior for non-structured meshes.
The main problem of shell finite elements with initially curved geometry is the so called ‘‘membrane lock-
ing’’. The EBST element has a quadratic interpolation of the geometry, then it may suffer from this prob-
lem. To assess this we resort to an example of inextensional bending. This is an hemispherical shell of radius
r = 10 and thickness h = 0.04 with an 18 hole in the pole and free at all boundaries, subjected to two
2428 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
1. 1
Normalized Displacement
RY
SY
ET
S YMM 1
MM
ET
RY
0. 9
DI
AP
H
B
RA
E 0. 8
FR E
GM
EBST-A
EBST-B
α =40 300 EBST1-A
EBST1-B
0
BST-A
0. 7
30
Z
BST-B
Y
X
(a) 1 2 3
10 10 10
(b) Number of DOFs
Fig. 7. Cylindrical roof under dead weight. E = 3 · 106, m = 0.0, thickness = 3.0, shell weight = 0.625 per unit area. (a) Geometry and
mesh for orientation B. (b) Displacement of point B for both (structured) mesh orientations.
Table 3
Cylindrical roof under dead weight
NDOFs Point-A Point-B
EBST EBST1 BST EBST EBST1 BST
16 0.65724 0.91855 0.74161 0.40950 0.70100 1.35230
56 0.53790 1.03331 0.74006 0.54859 1.00759 0.75590
208 0.89588 1.04374 0.88491 0.91612 1.02155 0.88269
800 0.99658 1.01391 0.96521 0.99263 1.00607 0.96393
3136 1.00142 1.00385 0.99105 0.99881 1.00102 0.98992
Normalized vertical displacements for mesh orientation A.
Table 4
Cylindrical roof under dead weight
NDOFs Point-A Point-B
EBST EBST1 BST EBST EBST1 BST
16 0.26029 0.83917 0.40416 0.52601 0.86133 0.89778
56 0.81274 1.10368 0.61642 0.67898 0.93931 0.68238
208 0.97651 1.04256 0.85010 0.93704 1.00255 0.86366
800 1.00085 1.01195 0.95626 0.99194 1.00211 0.95864
3136 1.00129 1.00337 0.98879 0.99828 1.00017 0.98848
Normalized vertical displacements for mesh orientation B.
Table 5
Cylindrical roof under dead weight
NDOFs Point-A Point-B
EBST EBST1 BST EBST EBST1 BST
851 0.97546 0.8581 0.97598 0.97662 1.0027 0.97194
3311 0.98729 0.9682 0.98968 0.98476 1.0083 0.98598
13,536 0.99582 0.9992 1.00057 0.99316 0.9973 0.99596
Normalized vertical displacements for non-structured meshes.
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2429
Free 1
Normalized Displacement
m. 0.8
Sy
Sy
m.
0.6
BST
0.4
EBST
QUAD
TRIA
F F 0.2
Free TRIC
EBST1
(a) 0
1000 2000 3000
(b) Number of nodes
Fig. 8. Pinched hemispherical shell with a hole: (a) geometry, (b) normalized displacement.
inward and two outward forces 90 apart. Material properties are E = 6.825 · 107 and m = 0.3. Fig. 8a
shows the discretized geometry (only one quarter of the geometry is considered due to symmetry).
In Fig. 8b the displacements of the points under the loads have been plotted versus the number of nodes
used in the discretization. Due to the orientation of the meshes chosen, the displacement of the point under
the inward load is not the same as the displacement under the outward load, so in the figure an average (the
absolute values) has been used. Results obtained with other elements have been included for comparison:
three membrane locking free elements, namely the original linear BST element, a transverse shear-deform-
able quadrilateral (QUAD) [32] and an assumed strain quadratic triangle (TRIC) [3]; a transverse shear
deformable quadratic triangle (TRIA) (standard displacement formulation for membrane part) [2] that is
vulnerable to locking.
From the plotted results it can be seen that the EBST element presents slight membrane locking in bending
dominated problems with initially curved geometries. This locking is much less severe than in a standard qua-
dratic triangle. Membrane locking disappears when only one integration point is used (EBST1 element).
The example is the inflation of a spherical shell under internal pressure. An incompressible Mooney–
Rivlin constitutive material has been considered. The Ogden parameters are N = 2, a1 = 2, l1 = 40,
a2 = 2, l2 = 20. Due to the simple geometry an analytical solution exists [33] (with c = R/R0):
h0 dW 8h0 6
p¼ 0 2
¼ 0 2
c 1 l1 l 2 c 2 :
R c dc Rc
In this numerical simulation the same geometric and material parameters used in [22] have been adopted:
R0 = 1 and h0 = 0.02. The three meshes of EBST1 element considered to evaluate convergence are shown in
Fig. 9a. The value of the actual radius as a function of the internal pressure is plotted in Fig. 9b for the
different meshes and is also compared with the analytical solution. It can be seen that with a few degrees
of freedom it is possible to obtain an excellent agreement for the range of strains considered. The final value
corresponds to a ratio of h/R = 0.00024.
The geometry of the dome and the material properties chosen are shown in Fig. 10. A uniform pressure
load of 600 psi is applied to the upper surface of the dome. The different meshes used in the analysis are
2430 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
Z Z Z
X Y Y Y
X X
(a)
4
3.5
2.5
Pressure
1.5
Exact
NDOFs=12
1
NDOFs=42
NDOFs=156
0.5
0
1 2 3 4 5
Radius
(b)
Fig. 9. Inflation of sphere of Mooney–Rivlin material. (a) Meshes of EBST1 elements used in the analysis, (b) radius as a function of
the internal pressure.
p h
ν=0.3
6 2
E=10.5 x 10 lb in
3 2
k y=24 x 10 lb in
R k' y=210 x 103 lb in2
-4
δ=2.45 x 10
α lb sec2 /in4
α = 26.67
o
R = 22.27 in.
h = 0.41 in.
Fig. 10. Spherical dome under impulse pressure. Geometry and material.
shown in Fig. 11. One fourth of the dome is considered only due to symmetry. Two different analyses under
elastic and elastic–plastic conditions were carried out. The number of thickness layers in Eq. (100) is four.
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2431
X Y
Fig. 11. Spherical dome under impulse pressure. Meshes used in the analysis. Mesh-1 with 338 elements, Mesh-2 with 1250 elements
and Mesh-3 with 2888 elements.
Numerical experiments show that this suffices to provide an accurate solution for large elastic–plastic prob-
lems [25]. Results are obtained using the explicit scheme.
Fig. 12 shows results for the time history of the central deflection using different meshes and elastic mate-
rial properties for both BST and EBST1 elements. Results are almost identical for mesh-2 and mesh-3,
showing the excellent convergence properties. The coarsest mesh shows some differences between both ele-
ments, but for the finer meshes the results are almost identical. Fig. 13 shows similar results but now for an
elastic–plastic material. The excellent convergence of the BST and EBST1 elements is again noticeable.
Results obtained with the present elements compare very well with published results using fine meshes.
See for example ABAQUS Explicit example problems manual [34] and WHAMS-3D manual [35], showing
plots comparing results using different shell elements.
A summary of results for the central deflection at significant times is given in Tables 6 and 7. Further
details on the solution of this problem with the standard BST element can be found in [25].
The geometry of the cylinder and the material properties are shown in Fig. 14. A prescribed initial
normal velocity of v0 = 5650 in./s is applied to the points in the region shown modelling the effect of
the detonation of an explosive layer. The panel is assumed clamped along all the boundary. One half of
0.05
BST mesh 3
EBST1 mesh 3
BST mesh 2
EBST1 mesh 2
BST mesh 1
Displacement [in]
0 EBST1 mesh 1
-0.05
-0.1
0 0.2 0.4 0.6 0.8 1
Time [msec]
Fig. 12. Spherical dome under impulse pressure. History of central deflection for elastic material.
2432 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
BST mesh 3
EBST1 mesh 3
BST mesh 2
-0.02 EBST1 mesh 2
BST mesh 1
Displacement [in]
EBST1 mesh 1
-0.04
-0.06
-0.08
0 0.2 0.4 0.6 0.8 1
Time [msec]
Fig. 13. Spherical dome under impulse pressure. History of central deflection for elastic–plastic material.
Table 6
Spherical dome: elastic material
Element/mesh t = 0.2 ms t = 0.4 ms t = 0.6 ms t = 0.8 ms
BST coarse 0.05155 0.09130 0.04414 0.08945
BST medium 0.04542 0.09177 0.03863 0.08052
BST fine 0.04460 0.09022 0.03514 0.08132
EBST1 coarse 0.05088 0.08929 0.04348 0.08708
EBST1 medium 0.04527 0.09134 0.03865 0.07979
EBST1 fine 0.04453 0.09004 0.03510 0.08099
Comparison of the central deflection values at the mid point obtained with the BST and EBST1 elements for different meshes.
Table 7
Spherical dome: elastic–plastic material
Element/mesh t = 0.2 ms t = 0.4 ms t = 0.6 ms t = 0.8 ms
BST Coarse 0.05888 0.05869 0.02938 0.06521
BST Medium 0.05376 0.06000 0.02564 0.06098
BST Fine 0.05312 0.05993 0.02464 0.06105
EBST1 Coarse 0.05827 0.05478 0.02792 0.06187
EBST1 Medium 0.05374 0.05884 0.02543 0.06080
EBST1 Fine 0.05317 0.05935 0.02458 0.06085
Comparison of the central deflection values at the mid point obtained with the BST and EBST1 elements for different meshes.
the cylinder is discretized only due to symmetry conditions. Three different meshes of 6 · 12, 12 · 32 and
18 · 48 triangles were used for the analysis. The deformed configurations for time = 1 ms are shown for
the three meshes in Fig. 15.
The analysis was performed assuming an elastic–perfect plastic material behaviour (ry ¼ k y k 0y ¼ 0). A
study of the convergence of the solution with the number of thickness layers showed again that four layers
suffice to capture accurately the non-linear material response [25].
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2433
3.0
h = 0.125 in 8
6 2
E = 10.5 x 10 lb/in
36 n w in
ec
v egio 03
in h
/s
ρ = 2.5 x 104 lb sec2/in4
50 i t
.2
10
ν = 0.33
12.56
0=
r
k y = 44 000 lb/in2
R=2.9375 k ’y = 0 lb/in2
Z Y
Fig. 14. Cylindrical panel under impulse loading. Geometry and material properties.
Y
X
Fig. 15. Impulsively loaded cylindrical panel. Deformed meshes for time = 1 ms.
A comparison of the results obtained with the BST and EBST1 elements using the coarse mesh and
the finer mesh is shown in Fig. 16 where experimental results reported in [36] have also been plotted for
1.4
1.2
1
Deflection [in]
0.8
0.6
0.4
Experimental
BST mesh 3
EBST1 mesh 3
0.2 BST mesh 1
EBST1 mesh 1
0
0 0.2 0.4 0.6 0.8 1
Time [msec.]
Fig. 16. Cylindrical panel under impulse loading. Time evolution of the displacement of two points along the crown line. Comparison
of results obtained with BST and EBST1 elements (mesh 1: 6 · 12 elements and mesh 3: 18 · 48 elements) and experimental values.
2434 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
Table 8
Cylindrical panel under impulse load
Element/mesh Vertical displacement (in.)
y = 6.28 in. y = 9.42 in.
BST (6 · 12 el.) 1.310 0.679
BST (18 · 48 el.) 1.181 0.587
EBST1 (6 · 12 el.) 1.147 0.575
EBST1 (18 · 48 el.) 1.171 0.584
Stolarski et al. [37] 1.183 0.530
Experimental [36] 1.280 0.700
Comparison of vertical displacement values of two central points for t = 0.4 ms.
comparison purposes. Good agreement between the numerical and experimental results is obtained. Fig. 16
show the time evolution of the vertical displacement of two reference points along the center line located at
y = 6.28 in. and y = 9.42 in., respectively. For the finer mesh results between both elements are almost iden-
tical. For the coarse mesh it can been seen again that the element BST is more flexible than element EBST1.
The numerical values of the vertical displacement at the two reference points obtained with the BST and
EBST1 elements after a time of 0.4 ms using the 16 · 32 mesh are compared in Table 8 with a numerical
solution obtained by Stolarski et al. [37] using a curved triangular shell element and the 16 · 32 mesh.
Experimental results reported in [36] are also given for comparison. It is interesting to note the reasonable
agreement of the results for y = 6.28 in. and the discrepancy of present and other published numerical solu-
tions with the experimental value for y = 9.42 in.
The deformed shapes of the transverse section for y = 6.28 in. and the longitudinal section for x = 0
obtained with the both elements for the coarse and the fine meshes after 1ms. are compared with the
experimental results in Figs. 17 and 18. Excellent agreement is observed for the fine mesh for both
elements.
2
Z [ in]
Original
Experimental
EBST1 mesh 1
1
BST mesh 1
EBST1 mesh 3
BST mesh 3
0 1 2
X [in]
Fig. 17. Cylindrical panel under impulse loading. Final deformation (t = 1 ms) of the panel at the cross-section y = 6.28 in.
Comparison with experimental values.
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2435
Z-direction [in]
2
Experimental
1 BST mesh 3
EBST1 mesh 3
BST mesh 1
EBST1 mesh 1
0
0 2 4 6 8 10 12
Y-direction [in]
Fig. 18. Cylindrical panel under impulse loading. Final deformation (t = 1 ms) of the panel at the crown line (x = 0.00 in.).
Comparison with experimental values.
Fig. 19. Inflation of a circular airbag. Deformed configurations for final pressure: (a) bending formulation; (b) membrane formulation.
2436 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
Fig. 20 shows the results obtained for the de-inflation process. Note that the spherical membrane falls
down due to the self weight. The configurations are, of course, non-unique.
q = 1000 kg/m3. Only one quarter of the geometry has been modelled due to symmetry. The thickness
h = 0.00075 m and the inflation pressure is 250,000 Pa. Pressure is linearly incremented from 0 to the
2438 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
final value in t = 0.15 s. The spherical object has a radius r = 0.08 m and a mass of 4.8 kg (one quarter)
and is subjected to gravity load during all the process.
The mesh has 8192 EBST1 elements and 4225 nodes on each surface of the airbag. Fig. 23 shows the
deformed configurations for three different times. The sequence on the left of the figure corresponds to
an analysis including full bending effects and that on the right is the result of a pure membrane analysis.
A standard penalty formulation is used in order to treat the frictionless contact.
The final problem corresponds to one of the sheet stamping benchmark tests proposed in NUMI-
SHEET96 [39]. The analysis comprises two parts, namely, simulation of the stamping of a S-rail sheet com-
Fig. 23. Inflation of a square airbag against an spherical object. Deformed configurations for different times. Left figure: results
obtained with the full bending formulation. Right figure: results obtained with a pure membrane solution.
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2439
Fig. 24. Stamping of a S-rail. Final deformation of the sheet after springback obtained in the simulation. The triangular mesh of the
deformed sheet is also shown.
ponent and springback computations once the stamping tools are removed. Fig. 24 shows the deformed
sheet after springback.
The detailed geometry and material data can be found in the proceedings of the conference [39] or in the
web [40]. The mesh used for the sheet has 6000 three node triangular elements and 3111 points (Fig. 24).
The tools are treated as rigid bodies. The meshes used for the sheet and the tools are those provided by the
benchmark organizers. The material considered here is a mild steel (IF) with Young modulus E = 2.06 GPa
and Poisson ratio m = 0.3. Mises yield criterion was used for plasticity behaviour with non-linear isotropic
hardening defined by ry(ep) = 545(0.13 + ep)0.267 [MPa]. A uniform friction of 0.15 was used for all the
tools. A low (10kN) blank holder force was considered in this simulation.
Fig. 25 compares the punch force during the stamping stage obtained with both BST and EBST1 ele-
ments for the simulation and experimental values. Also for reference the average values of the simulations
presented in the conference are included. Explicit and implicit simulations are considered as different
curves. There is a remarkable coincidence between the experimental values and the results obtained with
BST and EBST1 elements.
40
30
Punch Force [kN]
20
Experimental
Explicit (average)
10 Implicit (average)
BST
EBST1
0
0 10 20 30 40
Punch Travel [mm]
Fig. 25. Stamping of a S-rail. Punch force versus punch travel. Average of explicit and implicit results reported at the benchmark are
also shown.
2440 E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443
39
Z coordinate [mm]
38
37
Experimental
Explicit (average)
Implicit (average)
BST
EBST1
36
0 50 100 150 200 250
distance [mm]
Fig. 26. Stamping of a S-rail. Z-coordinate along line B00 –G00 after springback. Average of explicit and implicit results reported at the
benchmark are also shown.
Fig. 26 plots the Z coordinate along line B’’–G’’ after springback. The top surface of the sheet does not
remain plane due to some instabilities due to the low blank holder force used. Results obtained with the
simulations compare very well with the experimental values.
We have presented in the paper two alternative formulations for the rotation-free basic shell triangle
(BST) using an assumed strain approach. The simplest element of the family is based on an assumed con-
stant curvature field expressed in terms of the nodal deflections of a patch of four elements and a constant
membrane field computed from the standard linear interpolation of the displacements within each triangle.
An enhanced version of the BST element is obtained by using a quadratic interpolation of the geometry in
terms of the six patch nodes. This allows to compute an assumed linear membrane strain field which im-
proves the in-plane behaviour of the original element. A simple and economic version of the new EBST
element using a single integration point has been presented. The efficiency of the different rotation-free shell
triangles has been demonstrated in many examples of application including linear and non-linear analysis
of shells under static and dynamic loads, the inflation and de-inflation of membranes and a sheet stamping
problem.
The enhanced rotation-free basic shell triangle element with a single integration point (the EBST1
element) has proven to be an excellent candidate for solving practical engineering shell and mem-
brane problems involving complex geometry, dynamics, material non-linearity and frictional contact
conditions.
Acknowledgment
The problems analyzed with the explicit formulation were solved with the computer code STAMPACK
[41] where the rotation-free elements here presented have been implemented. The support of the company
QUANTECH (www.quantech.es) providing the code STAMPACK is gratefully acknowledged.
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2441
Appendix A
and
2 2 M 3 2 M 3 2 M 2 M 3 M 3 3
þLM
2;1 L2;1 þ L3;1 L3;1 þLM
2;2 L2;2 þ L3;2 L3;2 þLM
2;2 L2;1 þ L2;1 L2;2 þ L3;2 L3;1 þ L3;1 L3;2
6 LM L1 þ LM L3 LM 1 M 3
LM 1 M 1 M 3 M 3 7
6 1;1 3;1 3;1 2;1 1;2 L3;2 þ L3;2 L2;2 1;2 L3;1 þ L1;1 L3;2 þ L3;2 L2;1 þ L3;1 L2;2 7
6 M 1 7
6 L L þ LM L2 LM 1 M 2
LM 1 M 1 M 2 M 2 7
6 1;1 2;1 2;1 3;1 1;2 L2;2 þ L2;2 L3;2 1;2 L2;1 þ L1;1 Lj;3 þ L2;2 L3;1 þ L2;1 L3;2 7
BTb ¼ 6 7
6 M 1
L1;1 L1;1 LM 1
1;2 L1;2
M 1 M 1
L1;2 L1;1 þ L1;1 L1;3 7
6 7
6 2 2 M 2 M 2 7
4 LM2;1 L1;1 LM
2;2 L1;2 L2;2 L1;1 þ L2;1 L1;3 5
3 3 3 M 3
LM
3;1 L1;1 LM3;2 L1;2 LM3;2 L1;1 þ L3;1 L1;3
2 M 1 M 2 3
L1;1 q11 þ LM 2
1;2 q11 LM 1 M 2
1;1 q22 þ Li;2 q22 LM 1
1;1 q12 þ L1;2 q12
6 M 1 M 2 7
6 L2;1 q11 þ LM 2
2;2 q11 LM 1 M 2
2;1 q22 þ L2;2 q22 LM 1
2;1 q12 þ L2;2 q12 7
6 M 7
6 L q1 þ LM q2 LM q1 þ LM q2 LM 1 M 2 7
26 3;1 11 3;2 11 3;1 22 3;2 22 3;1 q12 þ L3;2 q12 7:
6 7
6 0 0 0 7
6 7
4 0 0 0 5
0 0 0
A.2. Membrane strain matrix and curvature matrix for the EBST element
References
[1] E. Oñate, A review of some finite element families for thick and thin plate and shell analysis, Publication CIMNE No. 53, May
1994.
[2] F.G. Flores, E. Oñate, F. Zárate, New assumed strain triangles for non-linear shell analysis, Comput. Mech. 17 (1995) 107–114.
[3] J.H. Argyris, M. Papadrakakis, C. Apostolopoulou, S. Koutsourelakis, The TRIC element. Theoretical and numerical
investigation, Comput. Methods Appl. Mech. Engrg. 182 (2000) 217–245.
[4] O.C. Zienkiewicz, R.L. Taylor, The finite element methodSolid Mechanics, vol. II, Butterworth-Heinemann, 2000.
[5] H. Stolarski, T. Belytschko, S.-H. Lee, A review of shell finite elements and corotational theories, Comput. Mech. Adv. 2 (2)
(1995).
[6] E. Ramm, W.A. Wall, Shells in advanced computational environment, in: J. Eberhardsteiner, H. Mang, F. Rammerstorfer (Eds.),
V World Congress on Computational Mechanics, Vienna, Austria, July 7–12, 2002. Available from: <http://wccm.tuwien.ac.at>.
[7] D. Bushnell, B.O. Almroth, Finite difference energy method for nonlinear shell analysis, J. Comput. Struct. 1 (1971) 361.
[8] S.P. Timoshenko, Theory of Plates and Shells, McGraw Hill, New York, 1971.
[9] A.C. Ugural, Stresses in Plates and Shells, McGraw Hill, New York, 1981.
E. Oñate, F.G. Flores / Comput. Methods Appl. Mech. Engrg. 194 (2005) 2406–2443 2443
[10] R.A. Nay, S. Utku, An alternative to the finite element method, Variational Methods Engrg. 1 (1972).
[11] J.K. Hampshire, B.H.V. Topping, H.C. Chan, Three node triangular elements with one degree of freedom per node, Engrg.
Comput. 9 (1992) 49–62.
[12] R. Phaal, C.R. Calladine, A simple class of finite elements for plate and shell problems. I: elements for beams and thin plates, Int.
J. Numer. Methods Engrg. 35 (1992) 955–977.
[13] R. Phaal, C.R. Calladine, A simple class of finite elements for plate and shell problems. II: an element for thin shells with only
translational degrees of freedom, Int. J. Numer. Methods Engrg. 35 (1992) 979–996.
[14] E. Oñate, M. Cervera, Derivation of thin plate bending elements with one degree of freedom per node, Engrg. Comput. 10 (1993)
553–561.
[15] M. Brunet, F. Sabourin, Prediction of necking and wrinkles with a simplified shell element in sheet forming, in: B. Kröplin (Ed.),
Int. Conf. of Metal Forming Simulation in Industry, vol. II, 1994, 27–48.
[16] G. Rio, B. Tathi, H. Laurent, A new efficient finite element model of shell with only three degrees of freedom per node.
Applications to industrial deep drawing test, in: M.J.M. Barata Marques (Ed.), Recent Developments in Sheet Metal Forming
Technology, 18th IDDRG Biennial Congress, Lisbon, 1994.
[17] J. Rojek, E. Oñate, Sheet springback analysis using a simple shell triangle with translational degrees of freedom only, Int. J.
Forming Processes 1 (3) (1998) 275–296.
[18] J. Rojek, E. Oñate, E. Postek, Application of explicit FE codes to simulation of sheet and bulk forming processes, J. Mater.
Process. Technol. 80–81 (1998) 620–627.
[19] J. Jovicevic, E. Oñate, Analysis of beams and shells using a rotation-free finite element-finite volume formulation, Monograph 43,
CIMNE, Barcelona, 1999.
[20] E. Oñate, F. Zárate, Rotation-free plate and shell triangles, Int. J. Numer. Methods Engrg. 47 (2000) 557–603.
[21] F. Cirak, M. Ortiz, Subdivision surfaces: a new paradigm for thin-shell finite element analysis, Int. J. Numer. Methods Engrg. 47
(2000) 2039–2072.
[22] F. Cirak, M. Ortiz, Fully C1-conforming subdivision elements for finite deformations thin-shell analysis, Int. J. Numer. Methods
Engrg. 51 (2001) 813–833.
[23] F.G. Flores, E. Oñate, A basic thin shell triangle with only translational DOFs for large strain plasticity, Int. J. Numer. Methods
Engrg. 51 (2001) 57–83.
[24] G. Engel, K. Garikipati, T.J.R. Hughes, M.G. Larson, L. Mazzei, R.L. Taylor, Continuous/discontinuous finite element
approximation of fourth-order elliptic problems in structural and continuum mechanics with applications to thin beams and
plates, and strain gradient elasticity, Comput. Methods Appl. Mech. Engrg. 191 (2002) 3669–3750.
[25] E. Oñate, P. Cendoya, J. Miquel, Nonlinear explicit dynamic analysis of shells using the BST rotation-free triangle, Engrg.
Comput. 19 (6) (2002) 662–706.
[26] F.G. Flores, E. Oñate, Improvements in the membrane behaviour of the three node rotation-free BST shell triangle using an
assumed strain approach, Comput. Methods Appl. Mech. Engrg. 194 (2005) 907–932.
[27] O.C. Zienkiewicz, E. Oñate, Finite elements vs. finite volumes. Is there really a choice?, in: P. Wriggers, R. Wagner (Eds.),
Nonlinear Computational Mechanics. State of the Art, Springer Verlag, Heidelberg, 1991.
[28] E. Oñate, M. Cervera, O.C. Zienkiewicz, A finite volume format for structural mechanics, Int. J. Numer. Methods Engrg. 37
(1994) 181–201.
[29] R. Hill, A theory of the yielding and plastic flow of anisotropic metals, Proc. R. Soc. London A 193 (1948) 281.
[30] R.W. Ogden, Large deformation isotropic elasticity: on the correlation of theory and experiments for incompressible rubberlike
solids, Proc. R. Soc. London A 326 (1972) 565–584.
[31] H.C. Huang, Static and Dynamic Analysis of Plates and Shells, Springer-Verlag, Berlin, 1989, p. 40.
[32] E.N. Dvorkin, K.J. Bathe, A continuum mechanics based four node shell element for general non-linear analysis, Engrg. Comput.
1 (1984) 77–88.
[33] A. Needleman, Inflation of spherical rubber ballons, Int. J. Solids Struct. 13 (1977) 409–421.
[34] Hibbit, Karlson and Sorensen Inc., ABAQUS, version 5.8, Pawtucket, USA, 1998.
[35] WHAMS-3D, An explicit 3D finite element program, KBS2 Inc., Willow Springs, Illinois 60480, USA.
[36] H.A. Balmer, E.A. Witmer, Theoretical experimental correlation of large dynamic and permanent deformation of impulsively
loaded simple structures, Air Force flight Dynamic Lab. Rep. FDQ-TDR-64-108, Wright-Patterson AFB, Ohio, USA, 1964.
[37] H. Stolarski, T. Belytschko, N. Carpenter, A simple triangular curved shell element, Engrg. Comput. 1 (1984) 210–218.
[38] P.O. Marklund, L. Nilsson, Simulation of airbag inflation processes using a coupled fluid structure approach, Comput. Mech. 29
(2002) 289–297.
[39] NUMISHEET96, Third International Conference and Workshop on Numerical Simulation of 3D Sheet Forming Processes,
NUMISHEET96, in: E.H. Lee, G.L. Kinzel, R.H. Wagoner (Eds.), Dearbon-Michigan, USA, 1996.
[40] Available from: <http://rclsgi.eng.ohio-state.edu/%7Elee-j-k/numisheet96/>.
[41] STAMPACK. A General Finite Element System for Sheet Stamping and Forming Problems, Quantech ATZ, Barcelona, Spain,
2005. Available from: <www.quantech.es>.