0% found this document useful (0 votes)
43 views19 pages

An Isogeometric Solid-Like Shell Element

An Isogeometric Solid-like Shell Element

Uploaded by

ficuni
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
43 views19 pages

An Isogeometric Solid-Like Shell Element

An Isogeometric Solid-like Shell Element

Uploaded by

ficuni
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 19

INTERNATIONAL JOURNAL FOR NUMERICAL METHODS IN ENGINEERING

Int. J. Numer. Meth. Engng 2013; 95:238–256


Published online 25 April 2013 in Wiley Online Library (wileyonlinelibrary.com). DOI: 10.1002/nme.4505

An isogeometric solid-like shell element for nonlinear analysis

Saman Hosseini1 , Joris J. C. Remmers1 , Clemens V. Verhoosel1 and René de Borst2, * ,†


1 Eindhoven University of Technology, Department of Mechanical Engineering, PO BOX 513, 5600 MB,
Eindhoven, The Netherlands
2 University of Glasgow, School of Engineering, Rankine Building, Oakfield Avenue, Glasgow G12 8LT, UK

SUMMARY
An isogeometric solid-like shell formulation is proposed in which B-spline basis functions are used to
construct the mid-surface of the shell. In combination with a linear Lagrange shape function in the thickness
direction, this yields a complete three-dimensional representation of the shell. The proposed shell element
is implemented in a standard finite element code using Bézier extraction. The formulation is verified using
different benchmark tests. Copyright © 2013 John Wiley & Sons, Ltd.

Received 28 October 2012; Revised 29 January 2013; Accepted 30 January 2013

KEY WORDS: shells; solid-like shell element; isogeometric analysis; Bézier extraction; geometrically
nonlinear analysis

1. INTRODUCTION

Isogeometric analysis (IGA) has recently received much attention in the computational mechanics
community. The basic idea is to use splines, which are the functions commonly used in computer-
aided design (CAD) to describe the geometry, as the basis function for the analysis rather than
the traditional Lagrange polynomial functions [1, 2]. Originally, non-uniform rational B-splines
(NURBS) have been used in IGA, but their inability to achieve local refinement has driven their
gradual replacement by T-splines [3].
A main advantage of IGA is that the functions used for the representation of the geometry are
employed directly for the analysis, thereby bypassing the need for a sometimes elaborate meshing
process. This important feature allows for a design-through-analysis procedure, which yields a
significant reduction of the time needed for preparation of the analysis model [2]. Indeed, the exact
parametrization of the geometry can have benefits for the numerical simulation of shell structures,
which can be very sensitive to imperfections in the geometry. Moreover, the higher-order continuity
of the shape functions used in IGA allows for a straightforward implementation of shell theories,
which require C 1 continuity such as Kirchhoff–Love models [4, 5]. A Reissner–Mindlin shell for-
mulation has been developed by Benson et al. [6] using NURBS basis functions. Although C 1
continuity is then no requisite, good results and a high degree of robustness were reported for large
deformation problems. In addition, the exact geometry description allows for an exact computation
of the shell director [7].
A further benefit of basis functions that possess a higher degree of continuity is that the com-
putation of stresses is vastly improved. In shell analysis, this can be particularly important when
materially nonlinear phenomena such as damage, or delamination, which can occur in laminated
spatial structures, are included in the analysis. In the latter case, the computation of an accurate

*Correspondence to: René de Borst, University of Glasgow, School of Engineering, Rankine Building, Oakfield Avenue,
Glasgow G12 8LT, UK.
† E-mail: [email protected]

Copyright © 2013 John Wiley & Sons, Ltd.


AN ISOGEOMETRIC SOLID-LIKE SHELL ELEMENT 239

three-dimensional stress field becomes mandatory, and solid-like shell elements become an obvious
choice [8–10]. The latter class of shell elements is characterized by the absence of rotational degrees
of freedom, which is convenient when stacking them, yet possess shell kinematics, and are rather
insensitive to shear locking and membrane locking.
Herein, we will develop a solid-like shell element that is based on the isogeometric concept.
It therefore combines the advantage of an accurate geometric description of the shell mid-surface
and the advantages in terms of meshing of IGA with the three-dimensional stress representation
of conventional solid-like shell elements. The formulation adopts NURBS (or T-spline) basis func-
tions for the discretization of the shell mid-surface, whereas in the thickness direction, conventional
Lagrange polynomials are used. The basic formulation is outlined in Section 2, whereas algorithmic
and implementation aspects are discussed in Section 3. Section 4 contains a number of benchmark
problems to assess the performance of the isogeometric solid-like shell formulation. So as not to
obscure the comparison by effects that are not purely caused by the shell formulation itself, no
allowance has been made for plasticity or damage, although locally the strains can be such that
these effects could occur. Furthermore, the comparison is restricted to monolayer shells, the exten-
sion to delamination being envisioned in a subsequent contribution. The paper concludes with some
observations on the efficiency of the use of basis functions with a high degree of smoothness for
problems, which exhibit highly localized deformation modes such as wrinkling.

2. SOLID-LIKE SHELL FORMULATION

As noted in the Introduction, shell elements that are based on the Reissner–Mindlin theory, and that
are commonly obtained by degenerating a solid, are less convenient for use in laminated spatial
structures, which is due to the presence of rotational degrees of freedom. With a view on extend-
ing the isogeometric shell element that we develop herein to laminated shell structures, we have
therefore taken the solid-like shell element proposed by Parisch [8] as point of departure, because
this element only possesses displacement degrees of freedom. Moreover, an internal stretch term is
added in this element, so that a quadratic term is obtained in the displacement field in the thickness
direction. As a consequence, the normal strain varies linearly, which significantly reduces mem-
brane locking and results in a full three-dimensional stress state. The original element by Parisch [8]
has been extended in [9, 10] for use in laminated composites, including interlaminar delamination.

2.1. Kinematics
Figure 1 shows the reference and the current configurations of the element, and the kinematics in
curvilinear coordinates. We describe the element kinematics through a linear combination of a pair
of material points at the top and at the bottom surfaces of the element. Each point at the top or at the
bottom surface of an element in the original configuration is labeled by its position vectors: Xt and
Xb , respectively. The variables  and  are the local curvilinear coordinates in the two independent
in-plane directions, and  is the local curvilinear coordinate in the thickness direction. The position

Figure 1. Geometry and kinematics of the shell in the undeformed and in the deformed configurations.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
240 S. HOSSEINI ET AL.

of a material point in the undeformed configuration is written as a function of the three curvilinear
coordinates:
X., , / D X0 ., / C D., / (1)
where X0 ., / is the projection of the point on the mid-surface of shell and D is the thickness
director at this point. In conventional solid-like shell elements, they are obtained as:
1
X0 ., / D ŒXt ., / C Xb ., /
2
(2)
1
D., / D ŒXt ., /  Xb ., /
2
In an isogeometric formulation, these quantities are computed directly, as we will show in the
remainder of this paper. The position of the material point in the deformed configuration x., /
is related to X., / via the displacement field ., , / as:
x., , / D X., , / C ., , / (3)
where
., , / D u0 ., / C u1 ., / C .1   2 /u2 ., / (4)
In this relation, u0 and u1 are the displacement of X0 on the shell mid-surface and the motion of
the thickness director D, respectively. The projection of the material point onto the mid-surface
leads to:
x0 ., / D X0 ., / C u0 ., / (5)
and:
d., / D D., / C u1 ., / (6)
Conventionally, the displacements u0 and u1 are calculated as:
1
u0 ., / D Œut ., / C ub ., / (7)
2

1
u1 ., / D Œut ., /  ub ., / (8)
2
which is convenient for applying the boundary conditions, although not strictly necessary in an iso-
geometric approach. In Equation (4), u2 is the internal stretching of the element, which is colinear
with the shell thickness director d in the deformed configuration. This quantity is expressed in terms
of stretch degree of freedom, w, through:
u2 D w., /ŒD., / C u1 ., / (9)
In any material point, a local reference triad can be established. The covariant base vectors are then
obtained as the partial derivatives of the position vectors with respect to the curvilinear coordinates
‚ D Œ, , . In the undeformed configuration, they are defined as:
@X
G˛ D D E˛ C D,˛ , ˛ D 1, 2
@‚˛ (10)
G3 D D
where ../,˛ denotes the partial derivative with respect to ‚˛ . E˛ is the covariant base vector defined
on the mid-surface:
@X0
E˛ D (11)
@‚˛

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
AN ISOGEOMETRIC SOLID-LIKE SHELL ELEMENT 241

Similarly, in the deformed configuration, we have:

@x
g˛ D D G˛ C u0,˛ C u1,˛ C .1   2 /u2,˛
@‚˛ (12)
g3 D D C u1  2u2

By using Equations (10) and (12), the components Gij and gij of the metric tensors in the
undeformed and the deformed configuration, respectively, can be determined as:

Gij D Gi  Gj , gij D gi  gj , i D 1, 2, 3 (13)

Elaboration of the expressions leads to the following components of the metric tensor in the
undeformed configuration:

G˛ˇ D E˛  Eˇ C ŒE˛  Dˇ C Eˇ  D˛ 
G˛3 D E˛ . D C D˛  D (14)
G33 D D  D

The elaboration of the components gij includes higher-order terms that are up to the fourth order in
the thickness coordinate  and derivatives of the stretch u2 with respect to the coordinates  and .
Neglecting these terms results in [8]:
  h i   h i h i  
g˛ˇ D E˛ C u0,˛  Eˇ C u0,ˇ C  E˛ C u0,˛  D,ˇ C u1,ˇ C Eˇ C u0,ˇ  D,˛ C u1,˛
      
g˛3 D E˛ C u0,˛  ŒD C u1  C  ŒD C u1   D,˛ C u1,˛  2u2  E˛ C u0,˛ (15)
g33 D ŒD C u1   ŒD C u1   4u2  ŒD C u1 

By using Equations (14) and (15), we can rewrite the components of the metric tensors as:

Gij D Gij0 C Gij1 , gij D gij0 C gij1 (16)

where Gij0 and gij0 correspond to the constant terms in Equations (14) and (15), whereas Gij1 and gij1
represent the linear terms. The contravariant base vectors needed for the calculation of the strains
can be derived as:

Gj D .G/1 Gi (17)

with G symbolically denoting the metric tensor in the undeformed configuration. The contravari-
ant base vectors denoted by the super script j are related to the corresponding base vectors on the
mid-surface via the so-called shell tensor jk as:
 
Gj D jk Ek , jk D ıkj   GN kj Ek (18)

where ıkj is the Kronecker delta and GN kj denotes the mixed variant metric tensor, which is calculated
with the covariant and contravariant tensors defined in Equation (16):

GN kj D G 0 j m G1 mk (19)

which is only nonzero when the undeformed shell is curved. The volume of the element in the
undeformed configuration is evaluated using the metric tensor Gij in the following manner:
p
d 0 D det .G/ d  d d  (20)

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
242 S. HOSSEINI ET AL.

2.2. Strain measure


The Green–Lagrange strain tensor  is defined conventionally in terms of deformation gradient F:
1 T 
D F FI (21)
2
where I is the unit tensor. The deformation gradient can be written in terms of the base vectors as:
F D gi ˝ Gi (22)
which leads to following representation of the Green–Lagrange strain tensor:
1
 D ij Gi ˝ Gj with ij D .gij  Gij / (23)
2
Substituting Equation (18) into this relation yields:
  
2 D .gij  Gij / ıki   GN ki ılj   GN lj Ek ˝ El (24)

Next, the strain tensor can be expressed in the membrane mid-surface strain,
ij , and the bending
strain, ij , as follows:

2 D .
ij C  ij /Ei ˝ Ej (25)
The strain components
ij and ij can be found in Appendix A.

2.3. Virtual work and linearization


In a Total Lagrangian formulation, the internal virtual work is expressed in the reference configura-
tion 0 :
Z
ıWint D ı T W S d0 (26)
0

with S the Second Piola–Kirchhoff stress tensor. The components of the virtual membrane strain,
ı
ij , and those of the virtual bending strain, ı ij , are given in Appendix B.
The resulting system of nonlinear equations is typically solved in an incremental-iterative man-
ner, which requires computation of the tangential stiffness matrix. This quantity is obtained by
linearizing the internal virtual work, Equation (26):
Z
 T   
D.ıWint / D ı W DS C D ı T W S d0 (27)
0

with ı and Dı defined in Appendix B.


In solid-like shell elements, the stresses are computed using a three-dimensional constitutive
relation. Assuming small strains, a linear relation between the rates of the Second Piola–Kirchhoff
stress tensor and the Green–Lagrange strain tensor can be adopted:
DS D C W D (28)
where C is the material tangential stiffness matrix. Substitution of this identity into Equation (27)
gives:
Z
 T   
D.ıWint / D ı W C W D C D ı T W S d0 (29)
0

3. ISOGEOMETRIC FINITE ELEMENT DISCRETIZATION

In this section, we review some basic concepts of IGA. Next, the Bézier extraction technique will be
outlined. This method is utilized to make a finite element data structure of the spline basis functions.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
AN ISOGEOMETRIC SOLID-LIKE SHELL ELEMENT 243

3.1. Fundamentals of B-splines and NURBS


A B-spline is a piecewise polynomial curve composed of linear combinations of B-spline basis
functions:
n
X
C./ D Ni ,p ./Pi (30)
i D1

where p is the order and n is the number of the basis functions. The Ni ,p ./ represents a B-spline
basis function and the coefficients Pi are points in space, referred to as control points. B-splines are
defined over a knot vector, „, which is a set of nondecreasing real numbers representing coordinates
in the parameter domain:

„ D ¹1 , 2 , : : : , nCpC1 º (31)

Parametric coordinates, i , divide the B-spline into sections. In one dimension, the B-spline basis
functions are defined recursively using the Cox-de Boor relation [11, 12]. By using tensor prod-
ucts, B-spline surfaces can be constructed using two knot vectors „ D ¹1 , 2 , : : : , nCpC1 º,
H D ¹1 , 2 , : : : , mCqC1 º and an n  m set of control points Pi ,j known as the control net. By
defining one-dimensional basis function Ni ,p and Mj ,p over these two knot vectors, the B-spline
surface is constructed as:
n X
X m
S., / D Ni ,p ./Mj ,q ./Pi ,j (32)
i D1 j D1

A drawback of B-splines is their inability to represent engineering objects such as conical sections
exactly. For this reason, NURBS, which encapsulate B-splines and can represent such objects
exactly, have become the standard in CAD. NURBS are defined by augmenting each control point
with a weight Wi > 0 as Pi D .xi , yi , ´i , Wi /. Such a point can be represented with homogeneous
coordinates Piw D .Wi xi , Wi yi , Wi ´i , Wi / in a projective R4 space. Accordingly, NURBS basis
functions are defined as:
N˛,p ./W˛
R˛,p D (33)
W./
P
where W./ D niD1 Ni ,p ./Wi is the weighting function. Note that there is no summation implied
over the repeated index ˛, and that a B-spline is recovered when all the weights are equal. The
NURBS surfaces are constructed by a tensor product of the one-dimensional functions as mentioned
in Equation (32), similar to B-splines.

3.2. Bézier extraction


As noted in the previous section, the parametric coordinates in a knot vector i divide the param-
eter domain into elements. Similar to the FEM, these elements, which refer to the knot intervals
¹i , i C1 º with a positive length, allow for piecewise integration using quadrature rules. On the other
hand, basis functions Ni ,p , have a local support over a knot interval ¹i , i CpC1 º, which means that
each element supports different basis functions, see Figure 2. This is at variance with the FEM where
numerical integration is done on a single parent element. To blend IGA into existing finite element
computer programs, we use Bézier elements and Bézier extraction operators [13] to provide a finite
element structure for B-splines, NURBS, and T-splines representations [14].
In general, a degree p Bézier curve is defined by a linear combination of p C 1 Bernstein basis
functions B./ [15]. Similar to the B-splines, by having an appropriate set of control points, a Bézier
curve is written as C./ D PT B. The Bézier extraction operator maps a piecewise Bernstein poly-
nomials basis onto a B-spline basis. This transformation makes it possible to use Bézier elements as
the finite element representation of B-splines, NURBS, or T-splines.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
244 S. HOSSEINI ET AL.

0.8

0.6

0.4

0.2

0
-1 0 1 -1 0 1 -1 0 1 -1 0 1

Figure 2. B-spline basis function plotted over Œ1, 1. The basis functions are different per element, which
is in contrast with standard finite elements.

The extraction operator can be obtained by means of knot insertion. Consider a knot vector „ and
a set of control points ¹Pk ºnkD1 . By inserting a knot value N in the knot vector, a new set of control
points needs to be calculated. This new set can be related to the initial set of control points via:
 T
PN D C1 P (34)

This relation ensures that the parametrization is not changed when an existing knot value is repeated,
see [13, 15] for algorithms to determine the operator C1 . The knot insertion process is repeated until
all interior knots of the knot vector have a multiplicity equal to p, with p the order of the original
spline defined over the knot vector „. Next, the complete set of new control points ¹PNk ºm kD1
, with
m D ne p C 1, is obtained as:
h iT h iT  T  1 T
N N
PN D CN N CN N 1    C2 C P D CT P (35)

Again, parametrization remains unchanged upon the insertion of the additional knots. Hence,
according to Equation (30) and by using Equation (35), it is expressed as:
 T
C./ D PT N./ D PN T B./ D CT P B./ (36)

Because P is arbitrary, the refined basis functions B are related to the original basis functions N via:

N./ D CB./ (37)

Hence, every original basis function can be expressed as a linear combination of the Bernstein poly-
nomials. By defining the operators Le and L N e to select the basis function Ne and Be , which are
defined over the elements, we have:

Ne D Le N D Le CLN e Be (38)

In this way, the element extraction operator can be elaborated as:


Ne
Ce D Le CL (39)

As it can be observed from Figure 3, the Bézier extraction operator of an element Ce maps a
piecewise Bernstein polynomial basis onto a B-spline basis.
The Bézier extraction operator for multivariate B-splines and NURBS can be computed by
exploiting their tensor product structure, see [13] for details. For a detailed discussion of Bézier
extraction for T-splines, for which a global tensor product structure is absent, see [14]. From
the element extraction operator, Bézier elements and the global Bézier mesh can be constructed,
see Figure 4.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
AN ISOGEOMETRIC SOLID-LIKE SHELL ELEMENT 245

1 1

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
-1 0 1 -1 0 1

Figure 3. Schematic representation of the Bézier extraction operator.

Figure 4. Schematic representation of Bézier extraction: (a) a Bézier element corresponding to a third-order
NURBS; (b) the Bézier element is mapped to the physical mesh using the Bézier extraction operator.

3.3. Isogeometric finite element implementation


The solid-like shell element as developed in [8–10] can formulated as an 8-node or as a 16-node
element. In both cases, a linear distribution of the internal stretch is assumed, so that only four
internal degrees of freedom, located at the four corners of the mid-surface of the element are neces-
sary. In this formulation, the projected displacements at the top and at the bottom surfaces, ut and
ub , are constructed using the degrees of freedom of the top and the bottom nodes, respectively. For
example, the displacement field for a quadrilateral 16-node element can be constructed using eight
biquadratic shape function used for the both top and the bottom nodes together with four bilinear
shape functions that are used for the discretization of the stretching.
We depart from the solid-like shell formulation outlined in Section 2, but we now only model the
mid-surface of the shell. Accordingly, the three-dimensional representation of the shell reduces to
a two-dimensional description using Bézier elements, where the geometric and the kinematic quan-
tities are approximated by NURBS functions. In a Bézier mesh, each control point Pi contains a
vector of degrees of freedom ˆ i , as follows:
 
ˆ i D axb , ayb , a´b , axt , ayt , a´t , aw T , i D 1, 2, : : : , np (40)

where ax , ay , a´ denote the displacement components and aw is the stretch degree of freedom. The
superscripts b and t correspond to the top and bottom surfaces of the shell, respectively, and np is
the number of control points in an element. For simplicity, we rewrite the vector ˆ i as:
 
ˆ i D aTi , aw T (41)

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
246 S. HOSSEINI ET AL.

 
where aTi D axb , ayb , a´b , axt , ayt , a´t corresponds to the bottom and the top degrees of freedom,
defined on the mid-surface of the shell. In a Bézier element, the stretching is interpolated at each
control point instead of varying linearly. To calculate the displacement terms in Equations (7) and
(8), we construct the matrix ‰ ei for a Bézier element e as:
2 3
Ni ,p 0 0
‰i D 4 0 Ni ,p 0 5 (42)
0 0 Ni ,p
with Ni ,p ., / the NURBS basis functions of order p. The displacements of the shell mid-
surface and the displacement of the thickness director, u0 and u1 , respectively, are obtained from
Equations (7) and (8), and can be discretized as:
u0 D N0 a, u1 D N1 a (43)
where N0 and N1 are defined as:
1
N0 D Œ‰ 1 , ‰ 1 , ‰ 2 , ‰ 2 ,    , ‰ np , ‰ np  (44)
2

1
N1 D Œ‰ 1 , ‰ 1 , ‰ 2 , ‰ 2    , ‰ np , ‰ np  (45)
2
and the vector a contains the translational degrees of freedom on the mid-surface:
h i
aT D aT1 , : : : , aTnp (46)

The stretch is interpolated with the same spline basis formulations as those for the in-plane
displacement field:
np
X
w., / D Ni ,p ., /.aw /i (47)
i D1

so that the interpolation matrix for the stretching can be written as:
Nw D ŒN1,p , N2,p ,    , Nnp ,p  (48)

The derivatives of the displacement vectors u0 and u1 with respect to the parametric coordinates
 and  follow in a standard manner as:
u0,˛ D N0,˛ a, u1,˛ D N1,˛ a (49)

with the matrices N0,˛ and N1,˛ containing the derivatives of the basis functions:
1
N0,˛ D Œ‰ 1,˛ , ‰ 1,˛ , ‰ 2,˛ , ‰ 2,˛    , ‰ np,˛ , ‰ np,˛  (50)
2
and
1
N1,˛ D Œ‰ 1,˛ , ‰ 1,˛ , ‰ 2,˛ , ‰ 2,˛    , ‰ np,˛ , ‰ np,˛  (51)
2
where:
2 @Ni ,p
3

0 0
6 7
‰ i ,˛ D 6
4 0
@Ni ,p

0 7
5 (52)
@Ni ,p
0 0 @˛

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
AN ISOGEOMETRIC SOLID-LIKE SHELL ELEMENT 247

3.4. Evaluation of internal force vectors and stiffness matrices


For the evaluation of the tangential stiffness matrices, we first define the virtual strain vector:

ı T D Œı 11 , ı 22 , ı 33 , 2ı 12 , 2ı 23 , 2ı 31  (53)

The virtual strain vector can be decomposed into components, as follow:

ı D H1 ıu1 C H01 ıu0,1 C H11 ıu1,1 C H02 ıu0,2 C H12 ıu1,2 C Hw ıw (54)

where the H matrices can be obtained from the strain variation, Appendix B. Next, the variations of
the displacement vectors are computed from the degrees of freedom associated to the control points
via the B-spline basis functions:

ıu1 D N1 ıa (55)

ıu0,˛ D N0,˛ ıa (56)

ıu1,˛ D N1,˛ ıa (57)

ıw D Nw ıw (58)

Substituting these relations into Equation (54) relates the vector of virtual strains to the control
points degrees of freedom:

ı D Bıˆ D Bu ıa C Bw ıw (59)

with the matrices Bu and Bw defined as:

Bu D H1 N1 C H01 N0,1 C H11 N1,1 C H02 N0,2 C H12 N1,2 (60)

Bw D Hw Nw (61)

Now, from the internal virtual work, Equation (26), the internal force vector is directly obtained as:
Z
fint D BT S d0 (62)
0

Next, we rewrite the linearized internal virtual work, Equation (29), in matrix form:

@fint
 D.ıWint / D ıˆ T Dˆ D ıˆ T K Dˆ D ıˆ T .Kmat C Kgeom / Dˆ (63)

where K represents the stiffness matrix decomposed in a material part Kmat and a geometric part,
Kgeom , as usual. From Equation (29), these matrices can be obtained as:
Z Z
mat T geom @BT
K D B CB d0 , K D S d0 (64)
0 0 @ˆ

The geometric part is the stress-dependent part of the stiffness matrix and is obtained through the
derivatives of the virtual strains, Appendix B.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
248 S. HOSSEINI ET AL.

3.5. Evaluation of the shell director


In the original solid-like shell formulation [8], the shell director D at the corners of an element were
derived from the positions of the corresponding nodes on the top and the bottom surface of this ele-
ment. The director in any arbitrary point on the mid-plane surface in the element followed from the
interpolation scheme. As a result, the shell director was at best a linear or quadratic approximation.
In the case of strongly curved shells, this may lead to significant errors.
In this paper, we calculate the shell director D directly by using the mid-surface base vectors
E˛ , ˛ D 1, 2, similar to [7] and identical to [5] (Chapter 15). The mid-surface vectors, which
are formulated in Equation (11), are discretized with a B-spline interpolation scheme. As a result,
these vectors represent exactly the in-plane directions of the shell and can be used to construct the
out-of-plane director D, according to:
E1  E2 t
DD . (65)
jjE1  E2 jj 2
In this relation, t is the thickness of the shell. Note that in the current implementation, it is assumed
that the thickness t is constant.

4. NUMERICAL SIMULATIONS

The isogeometric solid-like shell formulation is now verified and assessed through different bench-
mark tests. The NURBS-based CAD program Rhino has been used for modeling the geometry.
Because the present formulation is based on the discretization of the mid-surface of the shell, only
this surface has been modeled in Rhino. Subsequently, the Rhino model is converted to a data file
that contains the information on the control points, the elements, and the Bézier extraction operator.
This data file is exported to a code that works as the preprocessor, so we are able to select the control
points, elements, and edges for the application of the loads and other boundary conditions. A finite
element compatible input file is then generated, which is directly exported to the finite element code.
A 4  4 integration scheme has been used in all benchmark tests, see [16] for a detailed discussion
on integration schemes.

4.1. Clamped cylindrical shell


The first example concerns a clamped cylindrical shell subject to bending, Figure 5. The cylinder
has a radius R D 10 mm, a length L D 100 mm, a thickness t D 0.5 mm, a Young’s modulus
E D 2  107 N=mm2 , and a Poisson’s ratio D 0.3. It is clamped at one end and subjected to a ver-
tical load P at the other end. The exact geometry of the cylinder mid-surface has been constructed
using second-order or third-order NURBS basis functions.
The simulation has been carried out for various Bézier meshes, each consisting of M  N ele-
ments, where M and N are the number of elements in the longitudinal and the circumferential
directions, respectively. For the validation of this test, a linear analytical solution obtained from
beam theory is used, where the deflection of a beam ı classically reads PL3 =3EI .

Figure 5. Cylindrical shell modeled with four solid-like shell Bézier (SLSBEZ) elements.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
AN ISOGEOMETRIC SOLID-LIKE SHELL ELEMENT 249

Table I. Normalized deflection, p D 2.


Mesh 14 18 1  16 24 44
Deflection 0.8682 0.8711 0.8719 0.9901 1.032

Table II. Normalized deflection, p D 3.


Mesh 14 18 1  16 24 44
Deflection 1.012 1.017 1.036 1.028 1.034

Table III. Geometric parameters and material properties for the pinched
hemisphere.
Radius R Thickness t Young’s modulus E Poisson’s ratio
10.0 0.04 6.825  107 0.3

The results obtained for both orders of interpolation are shown in Tables I and II. In these tables,
the deflection is normalized with respect to the analytical solution obtained for a load P D 800 N.
Table I shows that, when using second-order basis functions, a response is obtained for the three
coarser meshes that is slightly stiffer than the analytical beam solution. Upon mesh refinement, an
increasingly softer response is computed, and for the finest mesh, the response is even softer than
the beam solution. Indeed, the kinematic assumptions that underlie the beam solution are slightly
different from the kinematics on which the present shell model is based, which explains that the
analytical solution is not retrieved exactly. This is even more pronounced when using third-order
basis functions. Then, all results are a bit softer than the beam solution, Table II.

4.2. Pinched hemispherical shell with hole


A pinched hemisphere with a hole at the top has been used extensively as a benchmark problem
for shell analysis to test the ability to describe nearly inextensional bending modes [17–19]. The
geometric parameters and the material properties employed in this test are summarized in Table III.
The shell is subjected to two opposite point loads. The bottom circumferential edge of the hemi-
sphere is free. Because of the symmetry, only a quarter of the shell needs to be modeled. The
symmetric boundary conditions are applied by constraining the displacement degrees of freedom
in the normal direction of the symmetry plane. The mesh and the applied boundary conditions are
shown in Figure 6. ABAQUS has been used to generate an alternative, traditional finite element
solution, using a 16  16 mesh consisting of so-called S4R shell elements.
Figure 7 shows the load-displacement curves of the pinched hemisphere that have been obtained
for different meshes. A plot of the deformed configuration and of the Von Mises stresses is given
in Figure 8. The results from a coarse mesh with 4  4 solid-like shell Bézier (SLSBEZ) elements
exhibit an overly stiff behavior. Increasing the number of elements to 8  8 results in a significant
softer response, whereas refining the mesh once more – to 16  16 elements – leads to results that
are close to the traditional finite element solution (using S4R elements). The graph also shows the
results from a 8  8 mesh of standard solid-like shell elements, which is between those obtained
with the medium mesh and with the fine mesh of SLSBEZ elements.
To make a first comparison between the results, we list the number of degrees of freedom for
the used meshes in Table IV. The number of degrees of freedom is not necessarily proportional
to the CPU time that is required to carry out a computation. For instance, the process of building
the stiffness matrix using Bézier extraction can be more time-consuming than for traditional finite
elements, but the smoother stress fields obtained using the SLSBEZ elements may result in a faster
convergence of the iterative process that is utilized to solve the nonlinear set of equations. Moreover,
the use of different finite element packages and different (linear) solvers to generate the results of

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
250 S. HOSSEINI ET AL.

free

sym sym

free

Figure 6. The mesh for a quarter model and the boundary conditions.

250
4× 4 SLSBEZ
8× 8 SLSBEZ
8× 8 SLS
200 16× 16 SLSBEZ
16 × 16 S4R

150
Load

100

50

0
0 0.5 1 1.5 2 2.5 3 3.5
Displacement

Figure 7. Diagram of the load P versus the displacement at point A for the pinched hemisphere.

Figure 8. Deformed configuration of the pinched hemisphere. Results for 16  16 SLSBEZ elements.

Table IV. Number of DOFs for different meshes employed in pinched hemispherical shell test.
Mesh 4  4 SLSBEZ 8  8 SLSBEZ 8  8 SLS 16  16 SLSBEZ 16  16 S4R
Number of DOFs 343 847 1606 2527 1445

Figure 7 make a direct comparison in terms of CPU time impossible. Nevertheless, it is clear that in
this example, elements that use splines as basis functions seem to be about as efficient as traditional
finite elements that employ Lagrangian polynomials. This is at variance with results reported for
other applications, and may be due to the fact that the C 0 continuity at the element boundaries of
traditional finite elements may actually facilitate the capturing of deformation patterns that exhibit
a locally strong curvature.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
AN ISOGEOMETRIC SOLID-LIKE SHELL ELEMENT 251

4.3. Pinched cylinder with free ends


The pinched cylinder with free ends shown in Figure 9 is used next to assess the element per-
formance. The parameters that define the geometry and the material properties are summarized
in Table V. The cylinder has free edges at the ends, and it is loaded by two centrally located
diametrically opposed point forces, which pull in the outward direction. Because of symmetry
considerations, only one-eight of the cylinder needs to be modeled.
The initial response is dominated by the bending stiffness, which induces large displacements
at relatively low load levels. This changes into a very stiff response when the displacement
becomes larger. Finite rotations occur afterwards, thus making the pinched cylinder with free ends
a challenging test for element performance [20–22].
Figure 10 shows the load-displacement curve obtained for different meshes of SLSBEZ elements,
for standard solid-like shell elements, and for the S4R shell elements implemented in ABAQUS. The
magnitude of the load is that for the complete cylinder and the displacement is measured at point A.
Figure 11 shows the deformed configuration and Von Mises stresses for a half model.
From Figure 10, it is inferred that all results are very close, with exception of the solution for the
8  8 mesh composed of SLSBEZ elements, which is slightly stiffer. Table VI gives the number of
degrees of freedom for each calculation and seems to indicate a somewhat better efficiency for finite
elements equipped with standard polynomials. This confirms the results obtained for the example
of the pinched hemisphere, obviously subject to the same reservations. It must be noted, however,
that efficiency is not the only criterion. For instance, the availability of a fully three-dimensional

Figure 9. Pinched cylinder with free ends.

Table V. Geometry parameters and material properties for the pinched cylinder
with free ends.
Radius R Length L Thickness t Young’s modulus E Poisson’s ratio
4.953 10.35 0.094 10, 500 0.3125

40
8× 8SLSBEZ
35 16× 16SLSBEZ
16× 8SLS
30 16 × 8S4R

25
Load

20

15

10

0
0 0.5 1 1.5 2 2.5 3
Displacement

Figure 10. Load-displacement diagram of pinched cylinder with free ends.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
252 S. HOSSEINI ET AL.

Figure 11. Deformed configuration of pinched cylinder meshed with 8  8 SLSBEZ elements.

Table VI. Number of DOFs for different meshes employed for the pinched
cylinder with free ends.
Mesh 8  8 SLSBEZ 16  16 SLSBEZ 16  8 SLS 16  8 S4R
Number of DOFs 847 2527 2598 765

stress field in solid-like shell elements makes this formulation richer and superior to standard shell
elements when material nonlinearities such as plasticity, damage, and delamination are included in
the analysis. Furthermore, a comparison that singles out a single displacement degree of freedom
may not be representative and obscures the full picture. For instance, the stress prediction using IGA
is vastly improved compared with standard finite elements, which has advantages, again particularly
for materially nonlinear analysis.

4.4. Pinched cylinder with rigid diaphragm


The problem of a pinched cylinder with a rigid diaphragm (see Figure 12) at the ends has been
studied by several authors [23–25] to test the convergence behavior and nonlinear performance of

Figure 12. Pinched cylinder with rigid diaphragm.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
AN ISOGEOMETRIC SOLID-LIKE SHELL ELEMENT 253

shell elements. Because large rotations occur, the problem provides a test for the finite rotation capa-
bility of the shell formulation. The dimensions and the material properties are shown in Table VII.
The cylinder is loaded by two centrally located, diametrically opposed point forces P , which push
inwards. By using symmetry, only one-eighth of the structure needs to be modeled.
Figure 13 shows results for a coarse, uniform mesh of 16  16 SLSBEZ elements and for a ref-
erence mesh of 40  40 S4R shell elements. The coarse mesh with SLSBEZ elements shows a far
too stiff behavior, which can be explained by the fact that it cannot well capture the effect of local
wrinkles. As a next step, results obtained through a local mesh refinement are presented in Figure 14.
As the wrinkles emerge close to the left edge of the structure, this part of the model is locally refined
using T-splines. The load-displacement curve shows an improvement, but it is not smooth and is still
stiffer than the reference shell solution. A second mesh refinement is carried out, which results in
the load-displacement curve shown in Figure 15, which is in good agreement with the reference
solution using S4R shell elements. To compare the performance of the SLSBEZ element with that
of the underlying solid-like shell element, we repeat the simulation by using a fine mesh of 80  80

Table VII. Geometry parameters and material properties for the pinched cylinder
with a rigid diaphragm at the ends.
Radius R Length L Thickness t Young’s modulus E Poisson’s ratio
100 200 1 30, 000 0.3

3000

2500

2000
Load

1500

1000

500
16× 16 SLSBEZ
40× 40 S4R
0
0 10 20 30 40 50 60
Displacement

Figure 13. Load-displacement diagram of pinched cylinder with rigid diaphragm. First mesh: 256 SLSBEZ
elements.

3000

2500

2000
Load

1500

1000

500
SLSBEZ,1threfinement
40× 40S4R
0
0 10 20 30 40 50 60
Displacement

Figure 14. Load-displacement diagram of pinched cylinder with rigid diaphragm. The right part of the figure
shows the first local mesh refinement using T-splines.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
254 S. HOSSEINI ET AL.

3000

2500

2000

Load
1500

1000

500 SLSBEZ, 2nd refinement


40× 40S4R
80× 80SLS
0
0 10 20 30 40 50 60
Displacement

Figure 15. Load-displacement diagram of pinched cylinder with rigid diaphragm. The right part of the figure
shows the second local mesh refinement using T-splines.

Table VIII. Number of DOFs for different meshes employed for the pinched
cylinder with a rigid diaphragm at the ends.
Mesh 847 SLSBEZ 80  80 SLS 40  40 S4R
Number of DOFs 6587 39, 366 8405

Figure 16. Deformed configuration of pinched cylinder with rigid diaphragm.

standard solid-like shell elements, see also Figure 15. The result clearly shows a stiffer response
than that of the finest mesh of SLSBEZ elements (see Table VIII).
Figure 15, finally, presents the deformed configuration. The plot of the Von Mises stresses shows
that locally very high stresses are computed, which normally will lead to plasticity or damage.
Material nonlinearities have not been included in the present analysis, however, so as not to mix the
results regarding the element performance with effects that can emanate from the integration of the
constitutive relation (also see Figure 16 for more details).

5. CONCLUDING REMARKS

A solid-like shell element has been formulated that is based on the isogeometric concept. Spline
basis functions (NURBS or T-splines) have been used to parametrize the mid-surface, whereas a
linear Lagrange shape function has been employed in the thickness direction. In this manner, a com-
plete three-dimensional representation of the shell is obtained. The shell formulation combines the
advantages of the solid-like shell formulation, such as the full, three-dimensional stress and strain
representation that allows for the straightforward implementation of constitutive relations such as

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
AN ISOGEOMETRIC SOLID-LIKE SHELL ELEMENT 255

plasticity or damage, with the advantages of IGA, including the exact description of the geometry,
the use of the design-through-analysis concept, and the accurate prediction of stress fields. The lat-
ter property is also beneficial for the prediction of the onset of plasticity, damage, or interlaminar
delaminations due to high transverse stresses.
In this study, the performance of the isogeometric solid-like shell element has been assessed by
means of a number of benchmark tests that involve geometric nonlinearity, comparing the new ele-
ment with conventional solid-like shell elements and with a standard shell element that is available
in a commercially available code. Although the spline-based solid-like shell element tends to per-
form slightly better than the conventional solid-like shell element, the observations regarding its
performance vis-à-vis the standard shell element are inconclusive, with the standard shell element
tending to perform better. It should be emphasized, however, that the comparisons are done, as
usual, on a global basis, that is, by comparing load-displacement curves. Different conclusions may
be reached if the predictions of the local stress fields are compared, or more generally, if a local prop-
erty is taken as benchmark property, which benefits from the higher smoothness of the spline basis
functions. Indeed, the analyzed examples suggest that the smoothness of spline functions can be
counterproductive in obtaining a fast convergence upon mesh refinement for problems that involve
highly localized deformations like local buckling, or wrinkling. This holds a fortiori when NURBS
are used, but is ameliorated through the use of T-splines, since they allow for an effective local
mesh refinement.
To keep the comparison as objective as possible, we have taken into account no materially non-
linear effects, although in certain cases, the stresses became so high that plasticity or damage would
have occurred. By ignoring these effects, however, the element performance was not diluted by
errors in the integration of the constitutive relation, nor by possible effects of a discontinuity, which
for instance enters at an elasto-plastic boundary.

APPENDIX A: STRAIN COMPONENTS

The strain field is derived from Equation (25) as:


2
˛ˇ D E˛  u0,ˇ C Eˇ  u0,˛ C u0,˛  u0,ˇ
2
˛3 D E˛  u1 C D  u0,˛ C u0,˛  u1
2
33 D 2D  u1 C u1  u1
2 ˛ˇ D Eˇ  u1,˛ C E˛  u1,ˇ C D,˛  u0,ˇ C D,ˇ  u0,˛ C u1,˛  u0,ˇ C u1,ˇ  u0,˛ (A.1)
h i  
 GN ˛ Eˇ  u0, C E  u0,ˇ C u0,  u0,ˇ  GN ˇ E˛  u0, C E  u0,˛ C u0,  u0,˛
2 ˛3 D D,˛  u1 C D  u1,˛ C u1  u1,˛
2 33 D 4u2  ŒD C u1 

APPENDIX B: VIRTUAL STRAINS AND DERIVATIVES

The virtual strain components are elaborated as:



˛ˇ D eˇ  ıu0,˛ C e˛  ıu0,ˇ

˛3 D d  ıu0,˛ C e˛  ıu1

33 D 2d  ıu1
2ı ˛ˇ D eˇ  ıu1,˛ C d,˛  ıu0,ˇ C e˛  ıu1,ˇ C d,ˇ  u0,˛ (B.1)
h i  
 GN ˛ eˇ  u0, C e  u0,ˇ  GN ˇ e˛  u0, C e  u0,˛
2ı ˛3 D d  ıu1,˛ C d,˛  ıu1
2ı 33 D 8u2  ıu1  4d  dıw
with ei D Ei C u0,i , i D 1, 2.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme
256 S. HOSSEINI ET AL.

The derivatives of the virtual strains, are given by:


2D.ı
˛ˇ / D ıu0,˛  Du0,ˇ C ıu0,ˇ  Du0,˛
2D.ı
˛3 / D ıu0,˛  Du1 C ıu1  Du0,˛
2D.ı
33 / D 2Du1  ıu1
2D.ı ˛ˇ / D ıu1,˛  Du0,ˇ C ıu0,ˇ  Du1,˛ C ıu1ˇ  Du0,˛ C ıu0,˛  Du1,ˇ (B.2)
h i  
 GN ˛ ıu0,  Du0,ˇ C ıu0,ˇ  Du0,  GN ˇ ıu0,˛  Du0, C ıu0,  Du0,˛
2D.ı ˛3 / D ıu1,˛  Du1 C ıu1  Du1,˛
 
2D.ı 33 / D 8 ıu1  dDw C ıwd  Du1 C wıu1  Du1

REFERENCES
1. Hughes TJR, Cottrell J, Bazilevs Y. Isogeometric analysis: CAD, finite element, NURBS, exact geometry and mesh
refinement. Computer Methods in Applied Mechanics and Engineering 2005; 194:4135–4195.
2. Cottrell J, Hughes TJR, Bazilevs Y. Isogeometric Analysis: Toward Integration of CAD and FEA. John Wiley &
Sons: Chichester, 2009.
3. Sederberg TW, Zheng J, Bakenov A, Nasri A. T-splines and T-NURCCs. ACM Transactions on Graphics 2003;
22:477–484.
4. Kiendl J, Bletzinger KU, Linhard J, Wüchner R. Isogeometric shell analysis with Kirchhoff–Love elements.
Computer Methods in Applied Mechanics and Engineering 2009; 198:3902–3914.
5. de Borst R, Crisfield MA, Remmers JJC, Verhoosel CV. Nonlinear Finite Element Analysis of Solids and Structures,
Second Edition. John Wiley & Sons: Chichester, 2012.
6. Benson DJ, Bazilevs Y, Hsu MC, Hughes TJR. Isogeometric shell analysis: the Reissner–Mindlin shell. Computer
Methods in Applied Mechanics and Engineering 2010; 199:276–289.
7. Dornisch W, Klinkel S, Simeon B. Isogeometric Reissner–Mindlin shell analysis with exact calculated director
vectors. Computer Methods in Applied Mechanics and Engineering 2013; 253:491-504. DOI: 10.1016/j.cma.2012.
09.010.
8. Parisch H. A continuum-based shell theory for non-linear application. International Journal for Numerical Methods
in Engineering 1995; 38:1855–1883.
9. Hashagen F, de Borst R. Numerical assessment of delamination in fibre metal laminates. Computer Methods in
Applied Mechanics and Engineering 2000; 185:141–159.
10. Remmers JJC, Wells GN, de Borst R. A solid-like shell element allowing for arbitrary delaminations. International
Journal for Numerical Methods in Engineering 2003; 58:2013–2040.
11. Cox MG. The numerical evaluation of B-splines. IMA Journal of Applied Mathematics 1972; 10:134–149.
12. de Boor C. On calculating with B-splines. Journal of Approximation Theory 1972; 6:50–62.
13. Borden MJ, Scott MA, Evans JA, Hughes TJR. Isogeometric finite element data structures based on Bézier extraction.
International Journal for Numerical Methods in Engineering 2011; 87:15–47.
14. Scott MA, Borden MJ, Verhoosel CV, Sederberg TW, Hughes TJR. Isogeometric finite element data structures based
on Bézier extraction of T-splines. International Journal for Numerical Methods in Engineering 2011; 88:126–156.
15. Piegl L, Tiller W. The NURBS Book, Second Edition. Springer-Verlag: New York, 1997.
16. Hughes TJR, Reali A, Sangalli G. Efficient quadrature for NURBS-based isogeometric analysis. Computer Methods
in Applied Mechanics and Engineering 2010; 199:301–313.
17. Betsch P, Gruttmann F, Stein E. A 4-node finite shell element for the implementation of general hyperelastic
3D-elasticity at finite strain. Computer Methods in Applied Mechanics and Engineering 1996; 130:57–79.
18. Buechter N, Ramm E. Shell theory versus degeneration – a comparison in large rotation finite element analysis.
International Journal for Numerical Methods in Engineering 1992; 34:39–59.
19. Simo JC, Fox DD, Rifai MS. On stress resultant geometrically exact shell model. Part III: computational aspects of
nonlinear theory. Computer Methods in Applied Mechanics and Engineering 1990; 79:21–70.
20. Gruttman F, Stein E, Wriggers P. Theory and numerics of thin elastic shells with finite rotations. Ingenieur-Archiv
1989; 59:54–67.
21. Sansour C, Bednarczyk H. The Cosserat surface as a shell model, theory and finite-element formulation. Computer
Methods in Applied Mechanics and Engineering 1995; 120:1–32.
22. Sansour C, Bufler H. An exact finite element rotations shell theory, its mixed variational formulation and its finite
element formulation. International Journal for Numerical Methods in Engineering 1992; 34:73–115.
23. Lindberg GM, Olson MD, Cowper GR. New development in the finite element analysis of shells. Quarterly Bulletin
of the Division of Mechanical Engineering 1969; 4:1–38.
24. Bucalem ML, Bathe KJ. Higher-order MITC general shell elements. International Journal for Numerical Methods
in Engineering 1993; 36:3729–3754.
25. Saleeb AF, Chang TY, Graf W, Yingyeunyong S. A hybrid/mixed model for non-linear shell analysis and its
application to large rotation problems. International Journal for Numerical Methods in Engineering; 29:407–446.

Copyright © 2013 John Wiley & Sons, Ltd. Int. J. Numer. Meth. Engng 2013; 95:238–256
DOI: 10.1002/nme

You might also like