You are on page 1of 42

Development of a non-linear triangular prism

solid-shell element using ANS and EAS techniques


Fernando G. Flores
Department of Structures, Universidad Nacional de Crdoba, Casilla de Correo 916,
5000 Crdoba-Argentina, and CONICET, fflores@efn.uncor.edu

Abstract
This paper extends a previous triangular prism solid element adequate to
model shells under large strains to become a solid-shell element, i.e. that
discretizations may include just one element across the thickness. A total
Lagrangian formulation is used based on a modified right Cauchy-Green de Three are the introduced modifications: a) an assumed
formation tensor (C).
mixed strain approach for transverse shear strains b) an assumed strain approach for the in-plane components using a four-element patch that includes
the adjacent elements and c) an enhanced assumed strain approach for the
through the thickness normal strain (with just one additional degree of freedom). One integration point is used in the triangle plane and as many as
necessary across the thickness. The intention is to use this element for the
simulation of shells avoiding transverse shear locking, improving the membrane behavior of the in-plane triangle, alleviate the Poisson effect locking
and to handle quasi-incompressible materials or materials with isochoric plastic flow. Several examples are presented that show the transverse shear and
Poisson effect locking free behavior, how the improvement in the membrane
approach alleviates the volumetric locking and the very good performance of
the introduced element for the analysis of shell structures for both geometric
and material non-linear behavior.
Keywords: Solid-Shell, Prism, Large strains
1. Introduction
Promoted by the continuous improvement in computer facilities and also
by the need for improving different aspects of the models to obtain more
reliable simulations, the use of solid elements for the simulation of shells
has grown a lot in the last fifteen years. The main advantages when using
solid-shell elements are: a) general tridimensional constitutive relations can
Preprint submitted to Elsevier

June 11, 2013

be used as plane stress state assumption is no needed; b) contact forces are


realistically applied on the outer surfaces which is particularly important for
friction; c) large transverse shear deformations can be considered specially
if more than one element is used across the thickness; d) special transition
elements between shell elements and solid elements are avoided; e) boundaries
non-parallel to the shell normal or director can be correctly modeled; f)
local triads and rotation vectors, that are costly in general and difficult to
parameterize and update, are not needed.
For the simulation of strongly non-linear problems due to, for instance,
complex constitutive models or contact, the simplex (linear interpolation)
elements are preferred and if possible with displacement degrees of freedom
(DOFs) only as they are more reliable and robust. Solid-shell elements, i.e.
when only one element is used across the thickness, developed until now
are hexahedral and particularly the largest developments are devoted to the
tri-linear 8-node brick. It is well known that simplex solid elements based
on the standard displacement formulation when used to simulate slender
structures lock severely. This numerical locking indicates the inability of
the interpolation functions (and their gradients) to fit the solid behavior
that many times makes the obtained solutions useless. In bending problems
transverse shear locking appears that increases with the shell slenderness.
Besides that a linear interpolation (constant gradient in that direction) does
not allow to fit a linear through the thickness strain due to the Poisson effect.
If the shell is initially curved artificial transverse strains and stresses appear
under pure bending due to curvature thickness locking. Membrane locking
specially appears on initially curved shells when bending is preponderant
without middle surface stretching. Finally when dealing with incompressible
or nearly incompressible materials or elastic-plastic materials with isochoric
plastic flow (metals typically) volumetric locking appears.
In order to be computationally efficient, it is necessary that solid-shell
elements have a different integration scheme on the shell tangent plane (as low
as possible) than across the thickness, where the number of integration points
must be arbitrary in order to capture the non-linearities of the constitutive
model when bending is present. This is particularly important for elasticplastic models for example springback in sheet metal forming simulations.
The use of a single integration point in the plane of the shell requires, in order
to maintain the efficiency, the series expansion of different variables (e.g.
the inverse of the Jacobian) and some control of the spurious deformation
modes due to under-integration. The first one brings some limitations on
the allowable distortion of the elements and the latter usually leads to the
introduction of user-defined factors that must be properly tuned. However
the advantages are very important and markedly decrease the storage space
2

for internal variables and the CPU time, particularly in codes with explicit
integration of the momentum equations.
The advances in solid and solid-shell elements aimed to cure the different
locking problems are numerous. In Reference [21] a detailed state of the art
for this type of elements can be found. To avoid transverse shear locking,
the classical mixed assumed strain approximation by Dvorkin and Bathe [6]
is the most used when full integration is considered (see for example [14, 29])
and a variation of it, modifying and increasing the sampling points to avoid
the occurrence of spurious modes, when reduced integration is preferred (see
for example [4]). In solid elements under large strains the volumetric locking has been mostly resolved using mixed elements with constant pressure
as originally proposed by [17], where the pressure degree of freedom can be
removed locally averaging over the element the volumetric strain, leading to
an element with constant volumetric strain. This technique is not acceptable for solid-shell elements because it leads to excessively flexible elements
and does not allow an adequate normal strain gradient across the thickness. More recently enhanced assumed strain techniques (EAS) [23] have
been developed consisting of improving the deformation gradient with the
inclusion of internal degrees of freedom which are condensed at the elementary level, maintaining displacements as the only global degrees of freedom.
EAS techniques not only allows to eliminate the volumetric locking, but also
to eliminate the problem due to the Poisson effect and can even be used
to improve the performance of the membrane part. It is not free of disadvantages: instabilities may occur under large compressive strain requiring a
significantly increase of elemental database or the use of an iterative loop in
each element for the determination of the internal degrees of freedom at each
global Newton-Raphson iteration. Apparently the solid-shell element with
the best performance developed until now is that presented in Refs. [20, 21]
where it is shown that an additional internal degree of freedom together with
a reduced integration scheme is sufficient to solve the volumetric locking and
the problem arising from Poisson effect. The element satisfies the membrane
and bending patch-tests, shows very good convergence properties and works
properly under large elastic-plastics strains as shown in [22]. Alternatively
[3] have developed an element with an additional local displacement degree
of freedom at the center of the element that improves the interpolation of
the normal displacement.
In the authors knowledge, the only triangular prism solid element adequate to simulate shell is the one proposed in [7]. In this Reference transverse
shear locking is cured using an assumed natural strain (ANS) to modify the
metric tensor components associated with shell normal direction. Volumetric
locking is alleviated on the one hand by averaging volumetric deformation
3

over the element (restricted to use at least two elements in the thickness)
and on the other hand by an assumed strain technique for the in-plane components that uses pieces of information from adjacent elements ([9]). Thus
a simple element is obtained, which does not require stabilization due to
under-integration, formulated in large strains and suitable for contact problems. One obvious and important advantage of a triangular prism element
is that the triangular mesh generators are quite more efficient, and provide
elements with a better aspect ratio. Curiously there are not developments
for triangular prism solid-shell elements. Probably the reason for this is the
few possibilities given by the interpolation functions of the standard prism
element.
The behavior of the standard (displacement base) 6-node prism and 8node brick are quite different, thats why the strategies to cure the different
locking problems may be different. The transverse shear locking of the first
one is quite lower while the latter has a better in-plane behavior. In a plane
strain condition a one point quadrature eliminates volumetric locking for a
quadrilateral but not for a triangle. Note also that for the same mesh density
(measured in the number of nodes) a reduced integration strategy implies the
double number of integration points for the prism than for the brick.
In this paper we propose to modify the assumed strain 6-node element
prism [7] in order to make it suitable as a solid-shell element, i.e., that just
one element can be used in the thickness of the sheet and get correct results
in thin shells with non-zero Poisson ratios and quasi-incompressible materials
or with isochoric plastic flow.
The next section summarizes the basic formulation of the solid element.
Then the improvements in the standard element are presented, starting with
the improvement in the tangent plane of the sheet, followed by the transverse
shear formulation and finally the EAS technique used to prevent the Poisson
effect locking. Section 5 presents several examples showing the very good
behavior of the element and finally some conclusions are summarized.
2. Basic kinematics of the solid finite element
Next the kinematic formulation of the standard 6-node triangular prism
element is presented. The element configurations are described by the standard isoparametric interpolations [31]
X () =
x () =

6
X
I=1
6
X

(1)

N I () XI
I

N () x =

I=1

6
X
I=1

N I () XI + uI

(2)

where XI , xI , are uI denote the original coordinates, the current coordinates


and the displacements of node I respectively. The shape functions N I ()
combine linear polynomials in terms of the area coordinates (, ) on the
triangular base with a linear interpolation () along the prism axis:
N 1 = zL1
N 2 = L1
N 3 = L1

N 4 = zL2
N 5 = L2
N 6 = L2

(3)

where the third triangular area coordinate and the linear Lagrangian polynomials have been included:
z = 1
1
(1 )
L1 =
2
1
L2 =
(1 + )
2

(4)

In a standard way, defining the Jacobian matrix at each integration point


J=

(5)

the computation of the Cartesian derivatives of the shape functions can be


performed
I
NX
= J1 NI
(6)
At each element center, coincident with the principal orthotropy directions of the constitutive material, a local Cartesian triad is defined
R = [t1 , t2 , t3 ]

(7)

that allows to compute the Cartesian derivatives with respect to this local
system (Y)
I
I
NY
= RT NX
(8)
and the deformation gradient F as a function of the present nodal coordinates
Fij =

NN
X

NyIj xIi

(9)

I=1

Finally the components of right Cauchy-Green deformation tensor C are


obtained
Cij = Fki Fkj
(10)
5

As most of the constitutive equations are easily dealt with using volumetric and deviatoric components, at each integration point the deformation
tensor is decomposed in a multiplicative form
1

C = [det (C)] 3 CD = J 3 CD

(11)

defining the volumetric and deviatoric strain components as


= ln J
(12)


1
2

eD = ln CD

the logarithmic strain tensor results

1 + eD
(13)
3
The computation of the strains (12) requires the spectral decomposition
of C
e=

C = L T 2 L

(14)

where 2 is a diagonal matrix collecting the eigenvalues 2i of C and L


includes the associated (unit) eigenvectors.
Adopting the hypothesis of additivity of elastic and plastic strain components, the strain tensor is written as
e = ep + ee

(15)

For materials with a yield surface independent of the mean press (Mises
or Hills yield functions for instance) it is possible to work exclusively with
the deviatoric strain an stress tensors, making easier and computationally
cheaper the integration of the constitutive equations.
As it will be shown in the next section the tensor C is modified using
assumed strain techniques that in one case includes an additional internal
The balance equations
degree of freedom , leading to an improved tensor C.
to be solved (variational formulation) are for the large strain case

g1 (u, ) =

Vo

g2 (u, ) =
Vo

1
S
2
1
S
2


: u C
dV0 + gext = 0
C

(16)


: C
dV0 = 0
C

(17)

where S is the second Piola-Kirchhoff stress tensor that can be related to


Kirchhoff stress tensor T using the following expressions:
a) defining the rotated tensor
TL = L T T L

(18)

b) the relationship between the rotated tensors is (See for example Reference
[5])
[SL ] =

1
[TL ]
2
(19)

[SL ] =

ln ( / )
 [TL ]
2 2

1
2

c) finally the 2nd Piola-Kirchhoff stress tensor is


S = L SL LT

(20)

As an alternative to the logarithmic strain, the spectral decomposition


(14) allows to easily deal with large strain hyperelastic materials (elastomers),
using models such as Ogden, Mooney-Rivlin, neo-Hookean, etc., that are
usually defined in terms of a strain energy written in terms of the principal
stretches.
3. Modifications of the standard element
The triangular prism element described above must be substantially improved to be used for large strain elastic-plastic analysis of shells including
contact constraints. Different modifications are introduced over the metric
tensor C with that aim.
The discretization of a shell type solid using solid-shell elements implies
two steps a) a discretization of the shell middle surface with three-node triangles in the present case and b) a discretization in the thickness direction
using one (or more) prism elements based on the triangles defined before.
Here it will be assumed that the 6-node connectivity associates nodes 1-3
and nodes 4-6 with planes nearly parallel to the shell middle surface and
that the latter ones are above the first three nodes along the shell normal
at a distance equal to the thickness. Thus middle surface normal direction
(local y3 ) is almost coincident with local natural coordinate .
As shown in previous section the strain tensor is computed from the
spectral decomposition of the right Cauchy-Green deformation tensor, thus
7

an interesting possibility is to directly modify the components of C associated


to the behavior intended to improve
m

m
s
C11 C12
C13
m
m
s
C22
C23
C = C21
(21)
s
s
C31 C32 C33
where the components with the upper index m are those that have the main
influence on the in-plane (membrane and bending) behavior of the shell and
those denoted with an s are those associated with the transverse shear. Then
the deformation tensor may be divided into three parts
(22)

C = C1 + C2 + C3
where


(23)



C2 = C13 t1 t3 + t3 t1 + C23 t2 t3 + t3 t2

(24)

C1 = C11 t1 t1 + C22 t2 t2 + C12 t1 t2 + t2 t1


corresponds with the components on the tangent plane,

are those components mainly associated with the transverse shear strains
and
C3 = C33 t3 t3

(25)

is used to compute the through the thickness strain.


The changes in each of the parts in which tensor C has been divided
are described next. The proposed approximation to C1 is identical to that
described in [7]. The modification in C2 is similar to that described in [7] but
modifying the position of the sampling points for the assumed mixed strain
approach. As for C3 now an EAS approximation is considered with the
inclusion of a single internal degree of freedom. Moreover unlike the original
reference that uses 2 fixed integration points in normal direction, here the
number of integration points is arbitrary, as usual in solid-shell elements.
For optimization reasons the integration scheme leads to a different way
of evaluating the equivalent nodal force vector and the geometric stiffness
matrix. With regard to the volume integral (performed with respect to the
reference configuration) the determinant of the Jacobian of the isoparametric
approach is evaluated at three points, namely the centers of the faces and the
element center, and is quadratically interpolated to the integration points.

..
.

P1

P2

P3

(a)

(b)

Figure 1: Patch of elements. (a) spatial view. (b) parametric space view.

3.1. Improvements on the in-plane behavior using the adjacent elements


To improve the in-plane interpolation the same technique used for rotationfree shell elements ([9]) is applied here. A four-element patch, involving 12
nodes, is defined by the element and its three adjacent ones (see Figure 1.a).
This allows to define an in-plane quadratic interpolation at both the upper
and lower surfaces, that is used to compute the in-plane deformation gradient and the associated components of the metric tensor. Here we follow the
procedure used in ([9]) that averages at the element center the metric tensor
components computed at each mid-side points. For the solid-shell element
exactly the same computations can be performed at both upper and lower
faces (see Figure 1.b with the notation of the lower face), and a subsequent
interpolation to the integration points. For the lower face the associated
quadratic shape functions are simply:
z
2

N 1 = (z + )

N7 =

N 2 = ( + z)

N 8 = 2 ( 1)

N 3 = ( + z)

N9 =

(z 1)
(26)

( 1)

Thus, at each face defined by three nodes of the element and another
three from the adjacent elements:
1. A local system (t1 , t2 ) on the shell tangent plane is computed (t1 and
t2 are chosen according to the local system R defined at the element
center), with t3 normal to the face .
9

2. At each mid-side point (PK ) the in-plane Jacobian (X , X ) is evaluated


and projected over the Cartesian directions (t1 , t2 )


X t1 X t1
J=
(27)
X t2 X t2
3. Note that at each mid-side point PK only four of the six derivatives of
the shape functions defined in (26)
1
2
3
7
8
9
1
1
I
N 1 + 1 z 2 z 2
0
1
I
N 1 + z 1 2 z
0
12

are non zero. For instance for mid-side point P1 (, ) = 12 , 21 the
shape function derivatives of nodes 8 and 9 that do not belong to any of
the two adjacent triangles are null. The Cartesian derivatives of these
four functions are computed as


N1I
N2I

K
=

J1
K

NI
NI

K

(28)

4. That allows to compute the in-plane deformation gradient (f1K , f2K ) and
with it CijK (i, j = 1, 2). These components are averaged over each face
f

C ij (f = 1, 2 for lower and upper face respectively).


5. When and adjacent element is missing (boundary), as originally proposed for rotation-free shells, the values of the components of Cij computed from the 3-node central triangle are included for the averaging.
For the prism element nG integration points are used along the normal direction (). At these points the in-plane components of the Cauchy-Green tensor
are interpolated using (remind that the modified components are identified
by an over bar)
Cij () = L1 Cij1 + L2 Cij2
(29)
while their variations (for

1
C
2 11
12 C22 =
C12

future use) are

1 1
C
2 11
1 1 1
C
L +
2 22
1
C12

1 2
C
2 11
1 2
C
2 22
2

C12

E11
L2 = E22
2E12

(30)

f relating the incremental tensor


At each face a modified tangent matrix B

10

components with the incremental


1 f

C
3
X
2 11
1
f
12 C22

=
3
f
K=1
C12

displacements u can be written as

1 K
C
2 11
1 K
C
2 22
K
C12

K J(K)
N
f
1
1
3
4
1 X X
J(K)
J(K)
f2K N2
=

 u
3 K=1 J=1
J(K)
J(K)
+ f2K N1
f1K N2

f
f
= B
m 318 u

(31)

where array uf include only the nodes on each face f (lower or upper).
Then it is possible to write

 
 1 1
2
, L2 B
m
B
L
(32)
B
=
m
m
336
Note that each matrix is associated with a different set of nodes, because
2 with the
1 is associated with the nodes on the lower face and B
matrix B
m
m
nodes on the upper face.
The equivalent nodal force vector stems from the integral

1 S11 T

 1 1
, L2 B
2 Jd u
S22
LB
rT1 u =
m
m
1
S12

S11

1 S11
1
1
1
2
2

S22 L Jd Bm u
S22 L Jd Bm ,
=

1 S12
S12

T
2 T
1

S11

S11
2
1
2
1

S22
Bm u
(33)
=
Bm , S22

2
1
S12

S12
where


uT = uT1 uT2 uT3 uT7 uT8 uT9 uT4 uT5 uT6 uT10 uT11 uT12

(34)

3.1.1. Geometric stiffness matrix


The geometric stiffness matrix results from
1
T

11

C
S
11
12
S22 udV
2 C22
(35)
uT Kg u =
V u

C12
S12


  J K
nG
2
3
4
X
 I I  S11 S12
VG X f X X
N1
I
L
u N1 N2
=
S21 S22
N2J
3
K=1 I,J=1
G=1
f =1
11

where the sum on G is the numerical integration with nG points along direction .
computed above (33)
It is possible to take advantage of the values S
leading to
T

u Kg u =

2 X
3
4
X
X

(
I

K
N1I N2I

G=1

f =1 K=1 I,J=1

(
2 X
3
4
X
X
f =1

"n

G
X
V G f S11
L
S21
3


K
u N1I N2I
I

f
S11
Sf
21

K=1 I,J=1

f
S12
Sf



22

)
K
N1J
J
u
N2J
)f
K

uJ
(36)

S12
S22
N1J
N2J

# 

where the contributions from each face may be seen independently one from
the other. In fact the integrated values Sijf depends on the strains at both
faces.
3.2. Transverse shear formulation
To cure transverse shear locking, following most of the literature, an interpolation in natural coordinates of mixed tensorial components is used.
In Reference [19] a general methodology for Reissner-Mindlin shell elements
is presented, that is particularized for the quadratic 6-node triangular element, leading to a linear variation of the transverse shear strain tangent to
the side. Assuming a constant value of the shear strain at each side, the
technique was also used for a linear 3-node element ([30]). For the present
element the latter case is helpful to write the relevant (mixed) components
of the right Cauchy-Green tensor as
1

 

2Ct3
C3
1
2
= P (, ) c
(37)
=
C3
C3
1

3
C3
with

1 1
1
2Ct3
2ft f3
2

c =
=
f2 f32
C3
3
f3 f33
C3

(38)

where the most relevant components to transverse shear (C3 , C3 ) has been
written with respect to a mixed coordinate system that includes the in-plane
natural coordinates (, ) and the spatial local coordinate in the transverse
direction (y3 ). These components are written in terms of the transverse shear

strain tangentto the side computed
at each mid-side point (1 = = 21 ,

2 = 0, = 21 and 3 = 21 , = 0 , see Figure 2.a). Similarly to hexahedral solid-shell elements with reduced integration (one point in the shell
plane)[4] the sampling points chosen are those located at lower and upper
12

faces ( = 1). As a result six sampling points located at mid-side of each


triangular face are necessary as shown in Figure 2.b. Besides that, the numerical integration is performed along the prism axis ( = = 13 ), whereby
1




  3 
2
3 
2Ct3 C3
+ C3
1
1
C3
C3
1
= P = , =
+
c = Pa c =
2
C3
+1
C3
3
3
3
(39)

ft

P2

P1

ft

f3

ft

f3

f3
3

P3
f

f2t 2

3
t

(a)

f3t

(b)

Figure 2: Points for the computation of the transverse shear strains.

Replacing (38) into (39) allows to compose




C2 = C3 t t3 + t3 t + C3 t t3 + t3 t
(40)
 3 
t t t
where the dualh base vectors
computed from the local base
i


X X
t t t3 = X
have been used. The modified Cartesian com y3
ponents (again denoted by an over bar) are computed projecting over the
local Cartesian base



C 13 = t1 C3 t t3 + t3 t + C3 t t3 + t3 t t3



(41)
= C3 a1 a33 + a31 a3 + C3 a1 a33 + a31 a3
where the symbol aji = ti tj (with i = 1, 2, 3 and j = , , 3) has been
introduced. Noting that aji = ij it simplifies to
C 13 = C3 a1 + C3 a1

(42)

Proceeding similarly for component C 23 we can write in a single expression



  



C 13
a1 a1
C3
C3
1
=
= Jp
(43)
C3
C3
C 23
a2 a2
13

where J1
p is the inverse of the in-plane Jacobian of the isoparametric mapping. Note that due to the way in which the local system has been defined
the components in (43) are null in the reference configuration.
At the sampling points the necessary deformation gradient components
are ft (natural coordinate derivative) and f3 (local Cartesian coordinate derivative). At sampling points ft are for the lower (1) and upper (2) faces respectively

1 3

x x2
2ft1
f2 = x1 x3
x2 x 1
f3

2 6

x x5
2ft1
f2 = x4 x6
x5 x4
f3

(44)

while f3 can be expressed as

f3 =

6
X

N3I xI =

f f f

I=1

y3

y3

y3

T
= (x) j3

(45)

where jT
are the components in direction y3 of the inverse of the Jacobian
3
of the isoparametric mapping (third column j1
3 ). Finally the modified transverse shear Cartesian components emerge from replacing equations (38) into
(39) and these into (43)
1 1


ft f3
C 13
1

(46)
= Jp Pa f2 f32
C 23
f3 f33
while the right Cauchy-Green tensor components at integration points are
obtained interpolating the values at each face
2




1
C13
C13
C13
1
L2
(47)
() =
L +
C23
C23
C23
Note that the Green-Lagrange strain components associated to the transverse shear at each face are directly the interpolated values




2E13
C 13
() =
()
2E23
C 23
s , relating displacement increments with strain inThe tangent matrix B
crements, is also obtained by interpolating from both faces
s () = B
1 L1 + B
2 L2
B
s
s
14

(48)

that requires first computing at the sampling points (for each face)

ft1 f31 + ft1 f31


s ue = f2 f32 f2 f32
B
f3 f33 + f3 f33
where (with xe = ue )
1 1 3

x x2
2ft
f2 = x1 x3
x2 x1
f3

(49)

2 6

x x5
2ft1
f2 = x4 x6
x5 x4
f3

(50)

and

f31 f32 f33

u1 u2 u3 u4 u5


6
u

1(1)

N3
2(1)
N3
3(1)
N3
4(1)
N3
5(1)
N3
6(1)
N3

1(2)

N3
2(2)
N3
3(2)
N3
4(2)
N3
5(2)
N3
6(2)
N3

1(3)

N3
2(3)
N3
3(3)
N3
4(3)
N3
5(3)
N3
6(3)
N3

(51)
then interpolate to the element axis as in (39) and finally convert to the
Cartesian system

f = Jf 1 Pa B
f
B
(52)
s
p
s
The associated nodal equivalent forces are computed as:
T

 
S13
T
s
r2 =
B
dV
218
S
23
V
1

 1 1
L +B
2 L2
=
QT B
Jd
s
s
218
1
1
1
T 1
1

2
=
Q L Jd Bs +
QT L2 Jd B
s
1
1
 1 

B
s
T
=Q
41
2
B
s
418

(53)

at each Gauss point are defined as


where the generalized shear forces Q



S
13
1
1
S23 L

Jd


Q41 =
(54)
S13

2
1
L
S23
15


3.2.1. Geometric stiffness matrix using Q
The above expressions allow to advance in obtaining the geometric stiffness matrix
 1  T

B
T
s

(55)
u KsG u =
2 u Q
B
s
where at each face we have
#
1 1 1
1 + f 2 f 2 + f 2 f 2 + 2f 3 f 3 + 2f 3 f 3

2f

2f

3
3
3
3
3

f u =
1t 13 1t
B
s
2ft f3 + 2ft f31 + 2f2 f32 + 2f2 f32 + f3 f33 + f3 f33
)
( 

1B
2 + 2B
3

B
s
s
s
u
(56)
= J1
p
1s 2B
2s + B
3s
3
B
J1
p
3

"

218

for instance
increments of strain components in the lower face
considering


1 u results
1, Q
2 B
Q
s

 J1
1, Q
2 p
Q
3

denoting by


2ft1 f31 2ft1 f31 + f2 f32 + f2 f32 + 2f3 f33 + 2f3 f33
2ft1 f31 + 2ft1 f31 + 2f2 f32 + 2f2 f32 + f3 f33 + f3 f33

 0 0 

1, Q
2 = Q
1, Q
2 J1
Q
p

(57)

we have


 1
1
1
1
01 + Q
02
2f

f
+
2f

Q
t
3
t
 1



3 +
1
0
0
2
2
2
2

Q1 , Q2 Bs u =
Q1 2Q2  f f3 f f3 +
3
01 + Q
02 f 3 f33 + f 3 f33
2Q

(58)
J(K)
related to f3 are
In the numerical implementation the derivatives N3
kept in an array but not those related to ft because its values are: 0, 1, and
-1 at each side.

3.3. Improvement of the transverse behavior


To avoid locking due to the Poisson effect when bending is important and
also to help in alleviating the volumetric locking (in quasi-incompressible
problems) an enhanced assumed strain formulation for the component C3 is
used.
3.3.1. Enhanced assumed strain (EAS) technique
The standard EAS method interpolates natural (convective) strain components. In this case we intend to improve directly the Cartesian component
C33 so a slightly different approach will be used as explained next.

16


At the element center = = 13 , = 0 the Cartesian deformation gradient component along y3 can be computed from the isoparametric interpolation
6
X
C
f3 =
N3IC xI
(59)
I=1

The enhanced gradient in direction y3 is defined as


f3 = f C e
3

(60)

thus the interesting component of the right Cauchy-Green tensor results in


C 2
C33 = f3C f3C e2 = C33
e

(61)

Note that for the standard element the normal strain in transverse directions results in a first approximation
p
p
C33 = f3 f3
(62)
3 =
1
e33 = ln (3 ) = ln (C33 )
(63)
2
that is practically constant in y3 direction for the linear element, whereas if
we now use the enhanced version

1
e33 =
ln C33
2

1
C 2
=
ln C33
e
2

1
C
ln C33
+
=
2
= eC
(64)
33 +
In this EAS approach the changes that will be produced by the enhanced
f3 on the other components C13 and C23 are disregarded, as they are computed
as explained in previous sections. Thus here only the influence on component
C33 is considered.
3.3.2. Balance equation and implicit solution
The variation of the Green strain now involves the internal DOF and
can be written
1
C33 = f3C f3C e2 + C33
E33 =
2
!
6
X
I
IC
=
N3 u f3C e2 + C33
=

I=1
2 C
e B3 ue

+ C33

17

(65)

In the last expression the first term replaces the corresponding part of the
standard displacement approach (the difference is the factor e2 on the components associated to E33 ).
The balance equation (17) associated to variable is
1
S33 C33 Jd = 0
(66)

thus must be solved iteratively. Denoting the residue by


1
S33 C33 Jd = r

(67)

the Newton-Raphson technique allows to approximate the value of nullifying the residue (67)

1
S33 C33
S33 C33
u +
Jd + r = 0
(68)
u

1
with
S33 C33
+ 2S33 e2 BC = C33 D3 B
+ 2S33 B
3
= C33 D3 B
3
u

S33 C33
= C33 D33 C33 + 2S33 C33 = C33 D33 C33 + 2S33

(69)
(70)

where D is the tangent constitutive matrix (Voigts notation ), D3 is its


is the matrix relating
third row (1 6) and D33 its diagonal component. B
incremental Green strains with incremental displacement when the assumed
3 is the first term of (65).
strain approximations are used for tensor C and B
When implicit techniques based on Newton-Raphson method are used to
obtain the equilibrium path, the DOF is locally condensed at element level
from equation (68)
1




+ 2S33 B
3 u + C33 D33 C33 + 2S33 Jd + r = 0
C33 D3 B
1

H118 u + k + r
then
=

1
r
Hu
k k
18

(71)
= 0
(72)

that is replaced in the corresponding balance equation associated to displacement variation (16)

T
T SdV uT Gext = uT r (u, )
B
(73)
u
V

once linearized

  T
 
1
T
S
S
B
B
T

B
u +
+
u +
S Jd
u

1
1
 T



DBu
T Jd
+ DT3 C33 + SGu + 2S33 B
B
3
1
1
 T



DB
+ SG u + B
T DT C33 + 2S33 B
T Jd
B
3
3
1

KT u + HT


1
r
T
KT u H
+ Hu
k k


r
T 1
H u HT
KT H
k
k

+r (u) = 0
+r (u) = 0
+r (u) = 0
+r (u) = 0
+r (u) = 0
+r (u) = 0
(74)

The contributions of the last expression to the Newton-Raphson scheme


are the modified elemental stiffness matrix and equivalent nodal force vector:
T = KT HT 1 H
K
k
r

r = r HT
k

(75)


C33 D33 C33 + 2S33 2 Jd
1
1
1
T

KT = KM + KG =
B DBJd +
SGJd
1
1
1

+ 2S33 B
3 Jd
H =
C33 D3 B

(77)

Note that above

(76)

k =

19

(78)
(79)

3.3.3. Geometric Stiffness Matrix


The contribution of the normal transverse component S33 to equilibrium
is

T
r3 =
B3 S33 dV
V
1
e2 BC
=
3 S33 Jd
=

BC
3 S33

where

(80)

S33 =

(81)

e2 S33 Jd
1

so the contribution to the geometric stiffness matrix is:


uT KG3 u = f3C f3C S33
=

6
X
I=1

6
 X
I T

N3IC N3JC S33

uJ

(82)

J=1

3.3.4. Explicit solution


A straightforward way to address the problem when using explicit integration of the momentum equations is to use the condition (72). The possible
steps are:
1. Modify the equivalent nodal forces
r = r + HT

r
k

(83)

and compute a first update of the EAS parameter

n+1 = n

r
k

(84)

2. Use the standard central difference scheme to compute the incremental


displacements u
3. Then compute a second update of the internal DOF
n+1 =
n+1

Hu
k

Such a scheme requires (besides keeping that is indispensable):


20

(85)

Keep k between steps, although it can be recomputed at the beginning


of the step the increment in the elemental data base is very low so it is
not worthwhile
Store row vector H to avoid its re-computation, this implies an important increase in the elemental data base
To know D3 (elastic-plastic relating increment in S33 (second PiolaKirchhoff) with increments in Green-Lagrange strains), that imply a
notably increase of floating point operations at each integration point.
This is because D3 is not available in explicit integrators (the same
problem appears for the stabilization of under-integrated elements)
To avoid the increase of the elemental data base [23] suggest (in implicit
codes) for models with a large number of internal DOFs (12) not to store the
corresponding submatrices (that here have been reduced to vector H and the
scalar k ) and to use an iterative scheme to update the internal DOFs in the
routine that compute stresses and equivalent nodal forces. These authors say
that with a few iterations (they suggest using just 2) accurate enough results
are obtained.
On the one hand, if the iterative process is reasonable accurate in a finite
element code with implicit integration of the momentum equations, it will be
even more accurate if the code is explicit. On the other hand each iteration
in the computation of the stresses is proportionally very costly in an explicit
code, so it will be assumed that just one iteration is enough to obtain accurate
results. The numerical experiments on elastic-plastic models indicate that
this assumption is correct even using the elastic component D33 (the iterative
process just requires the diagonal component not the entire row D3 ).
4. Numerical examples
In the set of examples shown below we denote by SPr the solid-shell element described above. A suffix Q indicates that the improvement in the membrane behavior has been activated. The original solid element on which the
solid-shell element is based is denoted by Prism (developed in [7]). With comparative purposes results obtained with other elements are includes: Q1SPs
is a reduced integration solid-shell hexahedral element with an excellent behavior [20, 21, 22] while LBST and BBST are rotation-free thin shell triangular
elements, the former [8] uses the standard constant strain triangle for the
membrane part, while the later includes an assumed strain approach for the
membrane part [10] almost identical to the formulation presented above.
These rotation-free shell elements compute bending strains resorting to the
21

Coordinates [mm]
1: (0.04, 0.02)
2: (0.18, 0.03)
3: (0.16, 0.08)
4: (0.08, 0.08)

3
4

2
1

Figure 3: Patch test

geometrical configuration of a four-element patch, the element and the three


adjacent ones, leading to a non-conforming approach (see [18] for a comprehensive treatment).
4.1. Patch test
The patch test is understood as a necessary condition for the convergence
of the element. In the case of solid elements it is expected that when nodal
displacements corresponding to a constant strain gradient (membrane patch
test) are imposed, constant efforts are obtained in all the elements. In the
case of standard shell elements displacements and rotations are imposed at
nodes corresponding to a constant curvature tensor (bending patch test) and
a constant moment tensor is expected to in all elements. In the case of a
solid-shell element clearly this must satisfy at least the membrane patch test
and, although it may not be necessary, it is highly desirable that the element
satisfies the bending patch test as this will lead to a more robust and reliable
element.
The Figure 3 shows a patch of elements that has been widely used to
access quadrilateral shell elements and hexahedral solid shell elements. Here
each hexahedral has been replaced by two prismatic elements. The size of the
largest sides is es a = 0.24mm and the size of the shortest side is b = 0.12mm,
while the thickness considered is t = 0.001mm. The lower surface has been
located at coordinate z = t/2. The mechanical properties of the material
are: Youngs modulus E = 106 MPa and Poisson ratio = 0.25. Because the
problem considered is linear just 2 integration points across the
thickness are
used located in the usual Gauss quadrature positions ( = 1/ 3).
4.1.1. Membrane patch test
The prescribed nodal displacements (on the boundary nodes) are defined
by the linear functions
y
ux = (x + ) 103
2

x
uy = (y + ) 103
2
22

(86)

and uz = 0 only on the nodes in the lower face to allow contraction due to
Poisson effect. Using present element SPr the correct results are obtained
for both the displacements of the interior nodes according to (86) and the
element stresses (xx = yy = 1333, 3 MPa y xy = 400 Mpa). The internal
DOF is null at all the elements. The same results are obtained with the
version SPrQ as the deformation gradient is constant.
4.1.2. Bending patch test
In this case the displacement field associated with a constant bending
stress state is given by (103 ):

y z
ux = x +
2 2


x z
uy = y +
2 2

uz = x2 + xy + y 2

1
2

(87)

that is prescribed on the exterior nodes of both shell faces. Again the results
obtained with both element versions are correct. The bending stresses at the
integration points are xx = yy = 0.3849 MPa y xy = 0.1155 MPa while
the displacements at the interior nodes correspond exactly with expression
(87). The internal DOF results = 0.3333 108 . Thus the element satisfy
the bending patch test also.
4.2. Cooks membrane problem
This example (see Figure 4) involves a large amount of shear energy
and is commonly used to assess in-plane bending performance. Plane strain
condition will be considered here with two different material behavior: a) a
quasi-incompressible elastic material with G = 80.1938GPa and K = 40.1
104 GPa corresponding with a Poisson ratio = 0.4999 and b) an elasticplastic material with elastic properties G = 80.1938GPa and K = 164.21GPa
implying a Poisson ratio = 0.29 and J2 plasticity with isotropic hardening
as a function of the effective plastic strain ep defined by
p

y = 0.45 + 0.12924ep + (0.715 0.45)(1 e16.93e ) [GPa].


The applied load is 100kN for the elastic case and 5kN for the elastic-plastic
material.
The plane strain condition implies coefficient C33 = 1 at all points ( =
0), thus the version without ANS for the in-plane components locks due to
the almost incompressibility constraint in the same way that a constant strain
triangle does. Because of that this example is intended to assess how the improvement in the membrane field collaborates to cure the volumetric locking.
The Figure 5 shows a convergence analysis as the mesh is refined. The vertical displacement of point C has been plotted versus the number of divisions
23

44

16

48

Figure 4: Cooks membrane problem. Geometry, units in mm.

per side. For comparison results obtained with three two-dimensional solid
formulation ([26])
elements have been included, a linear triangle with a F
that averages the volumetric component over two adjacent elements (only
results for the elastic case are available) and two 4-node quadrilaterals: Q1P0
entirely formulated in displacement with 2 2 integration for the shear components and the volumetric strain averaged over the element and Q1EA ([13])
based on the EAS technique including four internal degrees of freedom. In
the Figure 5.a the plots identified by 0.4999 and 0.499 indicates the Poisson
ratio used to obtained those results with element SPrQ. The first value is the
proposed Poisson ratio for the benchmark and the second a value slightly
lower that allows to assess the element sensitivity to the incompressibility
constraint. On one hand it can be seen that for the proposed Poisson ratio
both quadrilaterals show a clearly better behavior and that a large mesh
density is required to reach convergence with present element. On the other
hand the curve = 0.499 shows that in that case convergence is good enough
and similar to the triangle by [26]. For the elastic-plastic model (Figure 5.b)
where although plastic flow is isochoric the elastic behavior is compressible,
the results plotted indicate that present element has a better convergence
properties than both quadrilaterals..
4.3. Cantilever beam with a point load
This problem has been analyzed by a numerous of authors (see for example [25, 12, 21]). A cantilever plate strip of length L = 10mm width B = 1mm
and thickness t = 0.1mm is subjected to a transverse load F = 40N. For the
24

Vert. Displ. C [mm]

Vert. Displ. C [mm]

SouzaNeto
Q1PO
Q1E4
0.4999
0.499

10

15

20

25

30

Q1PO
Q1E4
SPrQ

35

Elements per side

10

15

20

25

30

35

Elements per side

(a)

(b)

Figure 5: Cooks membrane problem. a) elastic quasi-incompressible. b) J2 plasticity

selected Youngs modulus E = 106 MPa the behavior is one with large displacements but small strains. Using different values of Poisson ratio ( = 0.0,
= 0.3 , = 0.4999) it can be assessed if the proposed assumed strain techniques allow to avoid respectively the transverse shear locking, the Poisson
effect locking and the volumetric locking.
The final deformed configurations (vertical displacement is 70% of the
length) is achieved in ten equal load steps. The discretization includes 16
divisions in length, one in the width and one across the thickness with two
integration points. The Figure 6.a shows the vertical displacement of the tip
versus the load factor [0:1] for 5 different values of the Poisson ratio. The
case = 0 allows to compare with the reference value (uz = 7.08 mm) and
to see if the approach used to cure transverse shear locking is adequate. The
result obtained uz = 7.06 mm indicates that effectively the element is free
of transverse shear locking. The second value of Poisson ratio ( = 0.30)
is used to assess if the EAS technique avoids the appearance of locking due
to Poissons effect. In this case the computed displacement is uz = 7.01mm
that although is not exactly the same value obtained for = 0 shows that
the proposed method avoids the Poissons effect locking allowing a proper
gradation of the transverse normal strain. Finally the last three values of
Poisson ratio (0.49, 0.499 y 0.4999) allow to observe if the performance of the
element deteriorates significantly in the quasi-incompressible range. It can be
seen that although differences grow with Poisson ratio, this are below 4% for
the higher value considered. Besides, Figure 6.b plots the tip displacement as
25

1
7

Load Factor

0.8

0.6

Tip displacement [mm]

0.0
0.3
0.49
0.499
0.4999

6.75

0.4

6.25

0.2

0.0
0.3
0.499
0.4999
BBST 0.0

6.5

Tip Displacement [mm]

16

24

32

Number of elements

(a)

(b)

Figure 6: Bending of a cantilever plate strip

a function of the mesh density (number of divisions along the length) for four
different Poissons ratio. Results obtained with element BBST and = 0, that
for this example converges quite rapidly, are also plotted for comparison. For
the present element it can be seen that convergence deteriorates for Poissons
ratio larger than 0.499.
4.4. Cylindrical roof
This third example is the linear analysis of a cylindrical shell under self
weight. The roof is free along the straight sides and supported by rigid
diaphragms at the curved sides. Using symmetry considerations just onequarter of the roof is modeled. Five structured meshes were used to assess
convergence with the same number of elements along each side. Of the two
possible mesh orientations, the one shown in Figure 7.a was considered. As
the problem is linear just one element with two integration points is used
across the thickness.
As this is a membrane dominated problem, the membrane approximation is crucial for a fast convergence, then this example allows to assess the
importance of the ANS for in-plane components in non-isochoric problems.
The results for the vertical displacement at the mid-side point of the free side
(reference value is wA = 3.610.) are plotted in Figure 7.b as a function of
the number of elements per side. The figure includes two curves corresponding to both formulations of the membrane part of the solid-shell element.
Also the results obtained with shell elements LBST and BBST, which main
26

try
me
sym

sy
m
m
et
ry

3.5

di
ap
hr
ag
m

=40

free
300
Z

300

Displacement of A

LBST
BBST
SPr
SPrQ

2.5
2
1.5
1

0.5
0

10

15

20

25

30

35

Elements per side

(a)

(b)

Figure 7: Scordelis cylindrical roof. (a) geometry (b) displacement of point A

difference is the membrane approach, are included for comparison as their


behavior should be similar to the solid-shell element presented here. The results of present element SPr are almost identical to those obtained with the
thin shell element LBST except for the coarse mesh where the shell element is
more flexible because its bending approach is non-conforming. When using
the version with improved membrane behavior results quickly converge to
the reference value in a manner similar to the shell element with improved
membrane behavior BBST .
4.5. Semi-spherical shell with a 18o hole
The pinched hemisphere is considered in order to introduce initially double curved geometry. This is an extensively analyzed shell problem in the
context of large elastic displacements. The Figure 8.a shows the geometry
considered once symmetry conditions are introduced and the loads applied.
This is mainly an inextensional bending problem where the Poisson effect is
important and the membrane behavior is not. Because of the double curvature the curvature-thickness locking and the membrane locking may appear.
Two meshes have been considered that include 16 and 24 elements per side.
The coarse one is usually used to determine if the element suffers any locking
due to the initial curvature, i.e. if the results differs by more than 5% of the
target values some degree of locking exists. The middle surface radius is es
R = 10mm and the thickness is t = 0.04mm (R/t = 250). The mechanical
properties adopted are E = 6.825 104 GPa and = 0.3.
27

(a)

(b)

Figure 8: Semi-spherical shell with a hole. Original and deformed geometry.

The Figure 8.b shows the deformed configuration for an inward displacement of the loaded point equal to 60% of the shell radius. Whilst the Figure
9 plots the displacement (absolute values) of the loaded points where the
largest displacement corresponds to the inward load. The results presented
in Reference [24] using a shear deformable shell element (mesh 16 16) and
converged results obtained with element BBST using a 32 32 mesh are also
included. Results obtained with both formulations of present elements with
both meshes are plotted. Those obtained with the coarsest mesh (16 elements per side) are 8% lower than the target values, in contrast with those
presented in Reference [21] where an excellent approximation is obtained
with the same mesh using element Q1SPs. The results corresponding to the
finer mesh (24 elements per side) are in good agreement with the converged
results.
4.6. Slit annular plate
An annular plate with a radial cut is clamped at one of the slit edges
and subjected to a transverse load in the other one. The Figure 10.a shows
the original geometry and indicates the size and mechanical properties of the
material. The Figure 10.b shows the deformed configuration for the maximum load factor considered. This is a popular benchmark to assess shell
elements under large rotations. Initially proposed by [2] it has been considered by many authors and in Reference [27] converged results are presented
in tabular form that have been used here for comparison. The latter results have been obtained using the element SR4 present in commercial code
Abaqus[1] employing a mesh with 10 80 elements.
Two discretizations have been considered here, the first with 5 36 divisions and the second with 10 72 divisions. The vertical displacements
28

BBST
SIMO
SPrQ_16
SPrQ_24
SPR_16
SPR_24

Displacements

20

40

60

80

100

Load

(a)

(b)

Figure 9: Semi-spherical shell with a hole. Displacements of the loaded points.

B
Re=10

Ri=6
E=21x10

=0

thick. = 0.03

Pmax=0.8

Re
Ri

(a)

(b)

Figure 10: Slit Annular plate. (a) Initial geometry. (b) deformed geometry

29

Sze
SPr-5x36
SPrQ-5x36
SPrQ-10x72

0.8

P/P max

0.6

0.4

0.2

10

15

Displ. points A & B

Figure 11: Slit annular plate. Displacements of the free slit.

of two points on the loaded edge, denoted in Figure 10.b as A and B are
used for comparison. The Figure 11 plots the evolution of displacements
mentioned above in terms of load factor. It includes converged results from
Reference [27], results for both meshes considered when using SPrQ version
and results for the coarsest mesh for element SPr. A comparison of the results
for the coarse mesh allows to observe the influence of the ANS for membrane
part that for the maximum load factor indicate a difference in displacements
larger than 4%. Comparing the results obtained with the SPrQ version for
both meshes with the converged ones it can be seen that for low values of the
load factor the three curves almost coincide but for higher load factor the
displacements with the coarse mesh separate from the reference values until
almost a 4% for the maximum load factor. The results for the fine mesh are
in excellent agreement with those provided in [27]
4.7. Hinged cylindrical panel under point load
This example considers a rectangular cylindrical panel simple supported
along the straight sides and free along the curved sides, that is subjected
to a vertical point load in its center (see Figure 12). The middle surface
geometry is defined by the length of the panel L = 508mm, the radius of the
cylinder R = 2540mm and the half angle = 0.1rad. The behavior of the
panel presents a limit point, followed by a strong loss of strength and a final
stiffening once the curvature is inverted. Two different thicknesses for the
same mid-surface geometry have been considered t = 12.7 and t = 6.35 that
30

te
d
or
pp
Si

pl

su

Free

Z
Y
X

Figure 12: Cylindrical Panel under point load.

for the thin case leads to a snap back of the loaded point. This example has
been widely used to assess the performance of shell elements and non-linear
path-following techniques.
For this problem three meshes have been considered with 4, 6 and 8 elements per side. In this case 2 elements in the thickness direction have been
used that allows to introduce the hinge in the middle surface and then to
compare with solutions obtained with shell elements. The vertical displacement of the loaded point A (for both thicknesses) and the mid point of the
free side B (only for the thin case) are used to study convergence and for
comparison with other results. Two sets of converged results are included,
those obtained from reference [27] using a shear deformable shell element
(meshes with 16 16 and 2424 elements for the thick and thin case respectively) and our own results obtained with the thin shell element BBST using
a mesh with 16 16 elements. Also to compare with another solid-shell element results obtained for the coarsest mesh (4 4 2) with element Q1STs
[21] are plotted. For the thick case (Figure 13.a) there are some discrepancies when comparing the displacements obtained using present element with
shell elements but they are almost identical for the three meshes considered.
Also they are very similar to those obtained with the hexahedral solid-shell
element on the postcritical path. For the thin case (Figure 13.b) the results
obtained with the three meshes seem to converge to those obtained with shell
elements. The coarsest mesh is clearly inadequate to obtain reliable results
for the entire path, particularly for the descending post-critical path where
the snap back of the loaded point occurs. This can also be seem in the results
obtained with element Q1STs [21].
4.8. Square thin film under in-plane shear
This example has been previously analyzed in [28] using the commercial
code Abaqus[1]. Experimental data is also available [15]. The problem con31

3000

800

4el/s
6el/s
8el/s
Sze
Q1STs
BBST

2500

Point A
400

200

Load

Load

2000

Point B
600

1500

4el
6el
8el
Sze
Q1STs
BBST

1000
-200
500

-400

10

15

20

25

-600

30

Center Displacement

10

15

20

25

30

Displacement

(a)

(b)

Figure 13: Cylindrical panel. a) t = 12.7 b)t = 6.35.


=1

Figure 14: Square thin film under in-plane shear

sists of a square membrane (see Figure 14) with side a = 229mm made of
a thin film of Mylar with thickness t = 0.0762mm. The Mylar mechanical
properties are E = 3790MPa and = 0.38. The top and bottom edges
are clamped and the lateral edges are free. The top edge is subjected to a
uniform horizontal displacement = 1mm along the edge. The Figure 14
also show a perspective view of the deformed membrane (scaled 5X in the
transverse direction).
Three uniform structured mesh with 2626, 5151 and 101101 nodes,
with 1250, 5000 and 20000 elements respectively have been considered. In
the sequel these meshes will be referenced as mesh 25, 50 and 100, associated
to the number of subdivision along each side
Figure 15 plots two out-of-plane displacement profiles along the center of
the square in both Cartesian directions. The plot on the left corresponds to
32

25
50
100
BBST

Displ. z

Displ. z

25
50
100
BBST

-1
-1

50

100

150

200

Coord. x

50

100

150

200

Coord. y

(a)

(b)

Figure 15: Square membrane under in-plane shear. Transverse displacement profiles along
the center of the square. a) y = a/2. b) x = a/2.

y = 114.5 mm and the plot on the right to x = 114.5 mm. The results for the
three meshed defined above are included. These deformed configurations are
similar to the experimental evidence[15] and also coincide with the numerical
results presented in [28] obtained with program Abaqus [1] using the quadrilateral shell element S4R5. In that work mesh 100 was used and reported as
the minimum acceptable mesh according to their convergence studies. Here
the profiles for mesh 100 obtained with shell element BBST [11] have also
been included. For present element (SPrQ) the mesh 25 does not capture
the correct number of waves, but meshes 50 and 100 are almost coincident
on the wrinkled zones. It can be seen that the results are coincident with
those obtained with shell element BBST but on the free boundaries where the
membrane is in slack state. Note also that for mesh 50, that leads to very
good results, the element aspect ratio is x/t = 60.
4.9. Elastic-plastic conical shell
This a second example that allows to study the performance of present
element under large displacements and large elastic-plastic strains. The geometrical details are shown in Figure 16. The elastic mechanical properties
are E = 206.9MPa, = 0.29 whilst the plastic behavior obeys von Mises
yield function with isotropic hardening ruled by function
p
y (ep ) = 0.45 + 0.12924ep + 0.265 1 e16.93e
[MPa].
The reference load is P = 0.01N/mm that is scaled by the load factor .
33

R1= 1mm
R2= 2mm

R1

H = 1mm
t = 0.1mm
H

R2

Figure 16: Conical shell geometry

Due to the axisymmetric nature of the problem only a sector of = 0.1


rad is included in the discretization applying adequate kinematic constraints.
Four uniform discretization along the meridian direction has been considered
with 4, 8, 16 and 32 divisions. In the hoop direction only one division is
used and the number of integration points across the thickness has been set
to 5. The Figure 17 plots the load factor versus the vertical displacement of
the loaded line. The converged results obtained by [21] that discretize one
quarter of the geometry using 32 32 elements have been included. The
results obtained with element SPrQ for the four meshes are plotted but only
those for the coarsest mesh using element SPr, again to assess the influence of
the membrane improvement. From the plot it can be said that the mesh with
8 elements along the meridian gives results that are practically converged.
Element SPr is clearly stiffer than its counterpart SPrQ.
4.10. Deep drawing and elastic springback
This is one of the Numisheet93 benchmarks ([16]) where a U-shaped
(one axis bending) deep drawing is performed and the elastic springback is
measured. The Figure 18 shows the forming tools geometrical details. The
blank is made of an aluminum alloy with elastic properties E = 71GPa and
= 0.33. The plastic behavior is defined by the Lankford ratios r0 = 0.71,
r45 = 0.58 and r90 = 0.7 with isotropic hardening given by
y (ep ) = 576.79 (0.01658 + ep )0.3593 .
For the friction between blank and forming tools a coefficient = 0.162 has
been adopted while the blank folder force is 2.45 kN. The punch is first moved
70mm downwards and all the tools are then removed to allow the springback.

34

SPr-4
SPrQ-4
SPrQ-8
SPrQ-16
SPrQ-32
Q1STs

0.25

0.5

0.75

1.25

1.5

1.75

Displacement of the loaded line

Figure 17: Conical shell convergence study

Z
F/2

F/2
BLANK HOLDER

PUNCH
55

50

R5

55

R5

X
R5

70

STROKE

Load factor

DIE

52
Blank size: 350

Figure 18: Deep drawing of a strip

35

35

A
1

15

Figure 19: Elastic springback

This example assess the behavior of the element under moderate to large
plastic strains (less than 20%) and the bending behavior after such deformation. From symmetry consideration just one quarter of the problem has
been discretized using 10 uniform divisions in the half-width and 75 divisions in the half-length with mesh size of 5mm in the zones that are never
in contact with the the tools shoulder and mesh size of 1.5mm on the zones
that are plastically bent. The mesh size of 1.5 mm is the maximum that can
be used to correctly capture the contact due to the low radius of the tools
(5mm), as a higher mesh size leads on the one hand to erroneous contact
forces and on the other to a wrong prediction of the springback as discussed
in [3]. The Figure 19 shows a profile of the blank after the springback stage.
The measured parameters used for comparison with experimental values are
also shown in this figure. Using 5 integration point the following values have
been obtained (the values in brackets are the reported experimental range)
= 87mm [81 99], 1 = 111o [110 116] and 2 = 69o [68 76]. It can be
seen that all the parameters computed are within the experimental values.
4.11. Deep drawing of a square sheet
The last example considered is the deep drawing of a thin sheet corresponding to other of the benchmarks proposed in NUMISHEET93[16].
The Figure 20 shows the geometry of the tools. The undeformed sheet is
square with side length 150 mm. The elastic mechanical properties of the
mild steel considered are: elastic modulus E = 206GP a and Poisson ratio
36

85
35

Blank
Holder
Punch
35

Punch

Stroke
Min. 50

R10

35

R12

2
85

50
43
R5

48

Figure 20: Geometry of the tools (dimensions in mm) for the Numisheet 93 benchmark

= 0.3. For the plastic behavior the classical Hills yield function with
constant coefficients F = 0.283 , G = 0.358 , H = 0.642 , L = 1.065 ,
M = 1.179 , N = 1.289 was assumed. These coefficients were computed
from the Lankford ratios R0 = 1.79, R90 = 2.27 and R45 = 1.51. Isotropic
hardening is defined by the yield stress along rolling direction (X direction)
0 = 567.3 (0.00713 + ep )0.264 .
The symmetry conditions shown in the figure have been considered, then
just one quarter of the geometry has been included in the model. The discretization of the sheet includes 30 elements on each side (1800 elements in
the plane) and 7 integration points in the thickness direction. The blank
holder force used is 19,6 kN and the adopted friction coefficient is = 0.144.
The simulation considered a punch stroke of 40 mm.
The Figure 21 plots the punch force versus the punch travel. This figure
includes the results obtained with the 3-node triangular shell element BBST
with 7 through the thickness integration points and with the solid element
Prism with four layers, in both cases with the same in-plane mesh, for comparison. Results of element version SPr are not reported as in constrained
problem of this type it suffers from a severe volumetric locking due to the
isochoric plastic flow. It can be seen that the differences between the different
models are very small.
The Figure 22 shows the contour fills of the effective plastic strain for the
present solid-shell model (center), and those obtained with the solid (Prism)
and shell (BBST) models used for comparison. For the shell model the zone
with the largest plastic strain is less widespread and at the point with the
largest thickness increase (mid-side points) the equivalent plastic strain is
lower. The differences between the solid model and the solid-shell model are
37

80

Punch Force [kN]

60

40

SPrQ
NBST
Prism

20

10

20

30

40

Punch Travel [mm]

Figure 21: Punch force versus punch travel.

quite small.
5. Conclusions
In this paper we have developed a triangular prism solid-shell element
suitable for nonlinear analysis with elastic-plastic large strains. In the formulation assumed strain techniques have been used to prevent transverse
shear locking and to improve the membrane behavior. To avoid the locking due to the Poisson effect an enhanced assumed strain method with just
one internal degree of freedom has been proposed. The volumetric locking is
Z
Y

0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0

Prism-4 layers

SPrQ

BBST

Figure 22: Equivalent plastic strain for the final punch travel

38

alleviated as the combination of the effects of the membrane and the transverse approaches. The formulation is simple and can effectively achieve the
objectives. The main conclusions are:
Transverse shear locking disappears completely in all cases analyzed
The improvement in the membrane field is not only important in membrane dominated problems (e.g. deep drawing simulations), it results
crucial to cure the volumetric locking.
The EAS approach for the transverse normal strain effectively avoids
the locking due the Poisson effect in bending and collaborates in the
mitigation of the volumetric locking in quasi-incompressible problems.
For explicit time integration, one local (elemental) iteration for the
update of the internal degree of freedom seems to be enough to obtain
reliable results. Even for elastic-plastic problems the use of the elastic
mechanical parameters leads to the correct results.
For double curvature surfaced, the element converges to the correct results but the not with the same speed as reduced integration hexahedral
solid-shell elements.
The element did not show any problem in large elastic-plastic strain
cases.
Acknowledgments
The author acknowledges the financial support from CONICET (Argentina) and SeCyT-UNC.
[1] ABAQUS/Standard. Users Manual, version 6.3.1. Hibbit, Karlson and
Sorensen Inc., Pawtucket, EE.UU., 2002.
[2] Y. Basar and Y. Ding. Finite rotation shell elements for the analysis
of finite rotation shell problems. International Journal for Numerical
Methods in Engineering, 34:165169, 1992.
[3] B. Bassa, F. Sabourin, and M. Brunet. A new nine-node solid-shell finite
element using complete 3d constitutive laws. International Journal for
Numerical Methods in Engineering, 92, 2013.

39

[4] R.P.R. Cardoso, J.W. Yoon, M. Mahardika, S. Choudhry, R.J. Alves de


Sousa, and R.A. Fontes Valente. Enhanced assumed strain (eas) and assumed natural strain (ans) methods for one-point quadrature solid-shell
elements. International Journal for Numerical Methods in Engineering,
75:156187, 2008.
[5] M. Crisfield. Non-linear Finite Element Analysis of Solids and Structures II: Advanced Topics. John Wiley and Sons, 1997.
[6] E.N. Dvorkin and K.J. Bathe. A continuum based four-node shell element for general nonlinear analysis. Engineering Computations, 1:7788,
1984.
[7] F.G. Flores. A prism solid element for large strain shell analysis.
Computer Methods in Applied Mechanics and Engineering, 253:274296,
2013.
[8] F.G. Flores and E. Oate. A basic thin shell triangle with only translational dofs for large strain plasticity. International Journal for Numerical
Methods in Engineering, 51:5793, 2001.
[9] F.G. Flores and E. Oate. Improvements in the membrane behaviour of
the three node rotation-free bst shell triangle using an assumed strain
approach. Computer Methods in Applied Mechanics and Engineering,
194:907932, 2005.
[10] F.G. Flores and E. Oate. A rotation-free shell triangle for the analysis
of kinked and branching shells. International Journal for Numerical
Methods in Engineering, 69:15211551, 2007.
[11] F.G. Flores and E. Oate. Wrinkling and folding analysis of elastic
membranes using an enhanced rotation-free thin shell triangular element. Finite Elements in Analysis and Design, 47:982990, 2011.
[12] R.A. Fontes Valente, R.J. Alves de Sousa, and J.R.M. Natal. An enhanced strain 3d element for large deformation elastoplastic thin-shell
applications. Computational Mechanics, 34:3852, 2004.
[13] S. Glaser and F. Armero. On the formulation of enhanced strain finite
element methods in finite deformations. Engineering Computations, 14.
[14] R. Hauptmann and K. Schweizerhof. A systematic development of solidshell element formulations for linear and nonlinear analyses employing
only displacement degrees of freedom. International Journal for Numerical Methods in Engineering, 42:4970, 1998.
40

[15] J. Leifer, J.T. Black, Belvin W.K., and Behun V. Evaluation of shear
compliant boarders for wrinkle reduction in thin film membrane structures. 44th AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics and Materials Conference, 2003. Norfolk, Virginia, EEUU.
[16] A. Makinouchi, E. Nakamachi, E. Oate, and R. Wagoner. Proceedings
of the International Conference NUMISHEET93. 1993.
[17] J.C. Nagtegaal, D.M. Parks, and J.R. Rice. On numerically accurate
finite element solutions in the fully plastic range. Computer Methods in
Applied Mechanics and Engineering, 4:153177, 1974.
[18] E. Oate and F.G. Flores. Advances in the formulation of the rotationfree basic shell triangle. Computer Methods in Applied Mechanics and
Engineering, 194:24062443, 2005.
[19] E. Oate, O. Zienkiewicz, B. Surez, and RL. Taylor. A methodology for
deriving shear constrained reissner-mindlin plate elements. International
Journal for Numerical Methods in Engineering, 32:345367, 1992.
[20] M. Schwarze and S. Reese. A reduced integration solid-shell finite element based on the eas and the ans concept. geometrically linear problems. International Journal for Numerical Methods in Engineering,
80:13221355, 2009.
[21] M. Schwarze and S. Reese. A reduced integration solid-shell finite element based on the eas and the ans conceptlarge deformation problems.
International Journal for Numerical Methods in Engineering, 85:289
329, 2011.
[22] M. Schwarze, I.N. Vladimirov, and S. Reese. Sheet metal forming and
springback simulation by means of a new reduced integration solid-shell
finite element technology. Computer Methods in Applied Mechanics and
Engineering, 200:454476, 2011.
[23] J.C. Simo, F. Armero, and R.L. Taylor. Improved versions of assumed
enhanced strain tri-linear elements for 3d finite deformation problems.
Computer Methods in Applied Mechanics and Engineering, 110:359386,
1993.
[24] J.C. Simo, D.D. Fox, and M.S. Rifai. On a stress resultant geometrically
exact shell model. part iii: Computational aspects of the non-linear theory. Computer Methods in Applied Mechanics and Engineering, 79:21
70, 1990.
41

[25] J.C. Simo, M.S. Rifai, and D.D. Fox. On a stress resultant geometrically exact shell model. part iv: variable thickness shells with throughthe-thickness stretching. Computer Methods in Applied Mechanics and
Engineering, 81:91126, 1990.
[26] E.A. Souza Neto, F.M. Andrade Pires, and D.R.J. Owen. F-bar-based
linear triangles and tetrahedra for finite strain analysis of nearly incompressible solids. part i: formulation and benchmarking. International
Journal for Numerical Methods in Engineering, 62:353383, 2005.
[27] K.Y. Sze, X.H. Liu, and S.H. Lo. Popular benchmark problems for
geometric nonlinear analysis of shells. Finite Elements in Analysis and
Design, 40:15511569, 2004.
[28] A. Tessler, D.W. Sleight, and J.T. Wang. Effective modeling and nonlinear shell analysis of thin membranes exhibiting structural wrinkling.
AIAA Journal of Spacecrafts and Rockets, 42:287298, 2005.
[29] L. Vu-Quoc and X.G. Tan. Optimal solid shells for non-linear analyses of multilayer composites. i. statics. Computer Methods in Applied
Mechanics and Engineering, 192:9751016, 2003.
[30] F. Zrate, E. Oate, and F.G. Flores. A simple triangular element
for thick and thin plate and shell analysis. International Journal for
Numerical Methods in Engineering, 37:25692582, 1994.
[31] O.C. Zienkiewicz and R.L. Taylor. The finite element method. Vol II:
Solid Mechanics. Butterworth heinemann, 2000.

42

You might also like